Uploaded by Ishmam Zaman

Young & Freedman - University Physics with Modern Physics 14th ed US 2016 text

advertisement
SearS and ZemanSky’S
univeRSitY PHYSiCS
WitH moDeRn PHYSiCS
14tH eDition
HugH D. Young
RogeR A. FReeDmAn
University of California, Santa Barbara
ContRibuting AutHoR
A. LeWiS FoRD
Texas A&M University
Editor in Chief, Physical Sciences: Jeanne Zalesky
Executive Editor: Nancy Whilton
Project Manager: Beth Collins
Program Manager: Katie Conley
Executive Development Editor: Karen Karlin
Assistant Editor: Sarah Kaubisch
Rights and Permissions Project Manager: William Opaluch
Rights and Permissions Management: Lumina Datamatics
Development Manager: Cathy Murphy
Program and Project Management Team Lead: Kristen Flathman
Production Management: Cindy Johnson Publishing Services
Copyeditor: Carol Reitz
Compositor: Cenveo Publisher Services
Design Manager: Mark Ong
Interior and Cover Designer: Cadence Design Studio
Illustrators: Rolin Graphics, Inc.
Photo Permissions Management: Maya Melenchuk
and Eric Schrader
Photo Researcher: Eric Schrader
Senior Manufacturing Buyer: Maura Zaldivar-Garcia
Marketing Manager: Will Moore
Cover Photo Credit: Knut Bry
About the Cover Image: www.leonardobridgeproject.org
The Leonardo Bridge Project is a project to build functional interpretations of Leonardo da Vinci’s Golden Horn Bridge design,
conceived and built first in Norway by artist Vebjørn Sand as a global public art project, linking people and cultures in communities
in every continent.
Copyright ©2016, 2014, 2012 Pearson Education, Inc. All Rights Reserved. Printed in the United States of America. This publication is
protected by copyright, and permission should be obtained from the publisher prior to any prohibited reproduction, storage in a retrieval
system, or transmission in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise. For information
regarding permissions, request forms and the appropriate contacts within the Pearson Education Global Rights & Permissions department,
please visit www.pearsoned.com/permissions/.
Acknowledgements of third party content appear on page C-1, which constitutes an extension of this copyright page.
PEARSON, ALWAYS LEARNING and MasteringPhysics are exclusive trademarks in the U.S. and/or other countries owned by Pearson
Education, Inc. or its affiliates.
Unless otherwise indicated herein, any third-party trademarks that may appear in this work are the property of their respective owners
and any references to third-party trademarks, logos or other trade dress are for demonstrative or descriptive purposes only. Such references are not intended to imply any sponsorship, endorsement, authorization, or promotion of Pearson’s products by the owners of such
marks, or any relationship between the owner and Pearson Education, Inc. or its affiliates, authors, licensees or distributors.
CIP data is on file with the Library of Congress.
1 2 3 4 5 6 7 8 9 10—V303—18 17 16 15 14
www.pearsonhighered.com
ISBN 10: 0-321-97361-5; ISBN 13: 978-0-321-97361-0 (Student edition)
ISBN 10: 0-13-397798-6; ISBN 13: 978-0-13-397798-1 (BALC)
bRieF ContentS
MechAnics
1
2
3
4
5
6
7
8
9
10
11
12
13
14
Units, Physical Quantities, and Vectors
1
Motion Along a Straight Line
34
Motion in Two or Three Dimensions
67
Newton’s Laws of Motion
101
Applying Newton’s Laws
130
Work and Kinetic Energy
172
Potential Energy and Energy Conservation
203
Momentum, Impulse, and Collisions
237
Rotation of Rigid Bodies
273
Dynamics of Rotational Motion
303
Equilibrium and Elasticity
339
Fluid Mechanics
369
Gravitation
398
Periodic Motion
433
WAves/AcOustics
15 Mechanical Waves
16 Sound and Hearing
468
505
therMOdynAMics
17
18
19
20
Temperature and Heat
545
Thermal Properties of Matter
584
The First Law of Thermodynamics
618
The Second Law of Thermodynamics
647
electrOMAgnetisM
21
22
23
24
25
Electric Charge and Electric Field
683
Gauss’s Law
722
Electric Potential
752
Capacitance and Dielectrics
785
Current, Resistance,
and Electromotive Force
816
VOLUME 1: Chapters 1–20
• VOLUME 2: Chapters 21–37
26
27
28
29
30
31
32
Direct-Current Circuits
848
Magnetic Field and Magnetic Forces
881
Sources of Magnetic Field
921
Electromagnetic Induction
955
Inductance
990
Alternating Current
1020
Electromagnetic Waves
1050
Optics
33
34
35
36
The Nature and Propagation of Light
1078
Geometric Optics
1111
Interference
1160
Diffraction
1186
MOdern physics
37 Relativity
38 Photons: Light Waves Behaving
1218
39
40
41
42
43
44
Particles Behaving as Waves
1279
Quantum Mechanics I: Wave Functions
1321
Quantum Mechanics II: Atomic Structure
1360
Molecules and Condensed Matter
1407
Nuclear Physics
1440
Particle Physics and Cosmology
1481
as Particles
1254
Appendices
A
B
C
D
E
F
The International System of Units
Useful Mathematical Relations
The Greek Alphabet
Periodic Table of the Elements
Unit Conversion Factors
Numerical Constants
A-1
A-3
A-4
A-5
A-6
A-7
Answers to Odd-Numbered Problems
Credits
Index
A-9
C-1
I-1
• VOLUME 3: Chapters 37–44
The benchmark for clariTy and rigor
S
ince its first edition, University Physics has been renowned for its emphasis on fundamental
The challenge is to apply these simple conditions to specific problems.
principles and how
to apply them. This text is known for its clear and thorough narrative and for
Problem-Solving Strategy 11.1 is very similar to the suggestions given in Section 5.1
for the equilibrium of a particle. You should compare it with Problem-Solving
its uniquely broad,
deep,
and
thoughtful
set of worked examples—key tools for developing both
Strategy 10.1
(Section 10.2)
for rotational
dynamics problems.
conceptual understanding and problem-solving skills.
Problem-Solving Strategy 11.1 equilibrium of a rigid body
The Fourteenth Edition improves the defining features of the text while adding new features
influenced by physics education research. A focus on visual learning, new problem types, and
pedagogy informed by MasteringPhysics metadata headline the improvements designed to create the
best learning resource for today’s physics students.
344
Chapter 11 equilibrium and elasticity
for equilibrium 1g Fx = 0, gFy = 0, and gtz = 02 are applicable to any rigid body that is not accelerating in space and not
rotating.
any force whose line of action goes through the point you
choose. The body doesn’t actually have to be pivoted about an
axis through the reference point.
IdentIfy the relevant concepts: The first and second conditions
Set Up the problem using the following steps:
1. Sketch the physical situation and identify the body in equilibrium to be analyzed. Sketch the body accurately; do not represent it as a point. Include dimensions.
2. Draw a free-body diagram showing all forces acting on the
body. Show the point on the body at which each force acts.
3. Choose coordinate axes and specify their direction. Specify a
positive direction of rotation for torques. Represent forces in
terms of their components with respect to the chosen axes.
4. Choose a reference point about which to compute torques.
Choose wisely; you can eliminate from your torque equation
execUte the solution as follows:
1. Write equations expressing the equilibrium conditions. Remember
that gFx = 0, gFy = 0, and gtz = 0 are separate equations.
You can compute the torque of a force by finding the torque of
each of its components separately, each with its appropriate lever
arm and sign, and adding the results.
2. To obtain as many equations as you have unknowns, you may
need to compute torques with respect to two or more reference
points; choose them wisely, too.
a focus on Problem solving
locating your center of gravity WHile you WorK out
The plank (Fig. 11.8a) is a great way to strengthen abdominal,
back, and shoulder muscles. You can also use this exercise position to locate your center of gravity. Holding plank position with
a scale under his toes and another under his forearms, one athlete
measured that 66.0% of his weight was supported by his forearms
and 34.0% by his toes. (That is, the total normal forces on his
forearms and toes were 0.660w and 0.340w, respectively, where
w is the athlete’s weight.) He is 1.80 m tall, and in plank position
11.8 An athlete in plank position.
(a)
1.80 m
the distance from his toes to the middle of his forearms is 1.53 m.
How far from his toes is his center of gravity?
SolUtIon
IdentIfy and Set Up: We can use the two conditions for equilibrium, Eqs. (11.6), for an athlete at rest. So both the net force and
net torque on the athlete are zero. Figure 11.8b shows a free-body
diagram, including x- and y-axes and our convention that counterclockwise torques are positive. The weight w acts at the center
of gravity, which is between the two supports (as it must be; see
Section 11.2). Our target variable is the distance L cg , the lever arm
of the weight with respect to the toes T, so it is wise to take torques
78 (it
Chapter
with respect to T. The torque due to the weight is negative
tends 3 Motion in two or three Dimensions
to cause a clockwise rotation around T),
upward normal force at the forearms F is
problEm-solving stratEgy 3.1 projECtilE motion
a counterclockwise rotation around T).
gFx = 0 because there are no x
because 0.340w + 0.660w + 1- w2 = 0.
equation and solve for L cg :
execUte: The first condition for equilibrium
1.53 m
gtR = 0.340w102 - wL cg +
(b)
L cg = 1.01 m
evalUate: The center of gravity is
navel (as it is for most people) and closer
his toes, which is why his forearms
You can check our result by writing the
forearms F. You’ll find that his center of
forearms, or 11.53 m2 - 10.52 m2 =
M11_YOUN3610_14_SE_C11_339-368.indd 344
problem-solving strategies
coach students in how to approach
specific types of problems.
note: The strategies we used in Sections 2.4 and 2.5 for straighty
line, constant-acceleration
problems are also useful here.
identify the relevant concepts: The key concept is that through-
out projectile motion, the acceleration is downward and has a
constant magnitude g. Projectile-motion equations don’t apply to
throwing a ball, because during the throw the ball is acted on by
both the thrower’s hand and gravity. These equations apply only
after the ball leaves the thrower’s hand.
et uP the problem using the following steps:
1. Define your coordinate system and make a sketch showing
your axes. It’s almost always best to make the x-axis horizontal
and the y-axis vertical, and to choose the origin to be where
the body first becomes a projectile (for example, where a ball
leaves the thrower’s hand). Then the components of acceleration are ax = 0 and ay = - g, as in Eq. (3.13); the initial position is x0 = y0 = 0; and you can use Eqs. (3.19) through (3.22).
(If you choose a different origin or axes, you’ll have to modify
these equations.)
2. List the unknown
and known quantities, and decide which
9/11/14 11:48 AM
unknowns are your target variables. For example, you might
be given the initial velocity (either the components or the
magnitude and direction) and asked to find the coordinates
and velocity components at some later time. Make sure that
Discussion Questions
D
eVaLuate your answer: Do your results make sense? Do the
numerical values seem reasonable?
a body projECtEd horizontally
edge his velocity is horizontal, with magnitude 9.0 m>s. Find the
motorcycle’s position, distance from the edge of the cliff, and
velocity 0.50 s after it leaves the edge of the cliff.
at t = 0.50 s, we use Eqs. (3.19) and (3.20); we then find the distance from the origin using Eq. (3.23). Finally, we use Eqs. (3.21)
and (3.22) to find the velocity components at t = 0.50 s.
execute: From Eqs. (3.19) and (3.20), the motorcycle’s x- and
Tank
y-coordinates at t = 0.50
s are
bridging problems,
which
help
x = v t = 19.0 m>s210.50 s2 = 4.5 m
identify and set uP: Figure 3.22 shows our sketch of the trastudents
from
single-concept
jectory
motorcycle
and rider. He is in projectile
motion asmove
soon
Water height H
Waterofheight
y
Water
y = - gt = - 19.80 m>s 210.50 s2 = - 1.2 m
at t = 0
as he
leaves
at time
t the edge of the cliff, which we take to be the origin (so
x = y = 0). His initial velocity v at theworked
edge of the cliffexamples
is hori- The negative
to multi-concept
value of y shows that the motorcycle is below its
zontal (that is, a = 0), so its components are v = v cos a = starting point.
9.0 m>s and v = v sin a = 0. To find the
motorcycle’s position
Eq. of
(3.23),
the motorcycle’s
distance
from the origin at
d
problems
at theFrom
end
the
chapter,
have
t = 0.50 s is
been
revised,
based
on
reviewer
feedback,
3.22 Our sketch for this problem.
r = 2x + y = 214.5 m2 + 1- 1.2 m2 = 4.7 m
6. Use your result from step 5 to find the time when the tank is
At this point, the bike and
empty. How does your result depend on the initial height H?
From Eqs.
and (3.22), the velocity
ensuring that they
are(3.21)effective
andcomponents
at theat t = 0.50 s
rider become a projectile.
are
evaluate
7. Check whether your answers are reasonable. A good check is
appropriate
difficulty
v = vlevel.
= 9.0 m>s
to draw a graph of y versus t. According to your graph, what is
soLution
0x
2
1
2
solution Guide
0
identify and set uP
1. Draw a sketch of the situation that shows all of the relevant
dimensions.
2. List the unknown quantities, and decide which of these are the
target variables.
3. At what speed does water flow out of the bottom of the tank?
How is this related to the volume flow rate of water out of the
tank? How is the volume flow rate related to the rate of
change of y?
execute
4. Use your results from step 3 to write an equation for dy>dt.
5. Your result from step 4 is a relatively simple differential equation. With your knowledge of calculus, you can integrate it to
find y as a function of t. (Hint: Once you’ve done the integration, you’ll still have to do a little algebra.)
Problems
ExamplE 3.6
12.32 A water tank that is open at the top and
a hole atstunt rider rides off the edge of a cliff. Just at the
A has
motorcycle
the bottom.
execute the solution: Find the target variables using the equations you chose. Resist the temptation to break the trajectory into
segments and analyze each segment separately. You don’t have to
start all over when the projectile reaches its highest point! It’s
almost always easier to use the same axes and time scale throughout the problem. If you need numerical values, use g = 9.80 m>s2.
Remember that g is positive!
Solution
how long to drain?
A large cylindrical tank with diameter D is open to the air at the
top. The tank contains water to a height H. A small circular hole
with diameter d, where d V D, is then opened at the bottom of
the tank (Fig. 12.32). Ignore any effects of viscosity. (a) Find y, the
height of water in the tank a time t after the hole is opened, as a
function of t. (b) How long does it take to drain the tank completely?
(c) If you double height H, by what factor does the time to drain the
tank increase?
you have as many equations as there are target variables to be
found. In addition to Eqs. (3.19) through (3.22), Eqs. (3.23)
through (3.26) may be useful.
3. State the problem in words and then translate those words into
symbols. For example, when does the particle arrive at a certain
point? (That is, at what value of t?) Where is the particle when
its velocity has a certain value? (That is, what are the values of
x and y when vx or vy has the specified value?) Since vy = 0
at the highest point in a trajectory, the question “When does
the projectile reach its highest point?” translates into “What is
the value of t when vy = 0?” Similarly, “When does the projectile return to its initial elevation?” translates into “What is
the value of t when y = y 0?”
389
Solution
bridging problEm
A research-based problem-solving approach—
identify, set Up, execUte, evalUate—is used
in every Example and throughout the Student’s and
Instructor’s Solutions Manuals and the Study Guide. This
consistent approach teaches students to tackle problems
thoughtfully rather than cutting straight to the math.
Solution
examPle 11.2
evalUate your answer: Check your results by writing gtz = 0
with respect to a different reference point. You should get the
same answers.
S
0
0
0y
0x
0
0
2
0
x
For assigned homework and other learning materials, go to MasteringPhysics®.
2
0
2
the algebraic sign of dy>dt at different times? Does this make
sense?
1
2
0
2
2
2
0x
vy = - gt = 1 - 9.80 m>s2210.50 s2 = - 4.9 m>s
The motorcycle has the same horizontal velocity vx as when it left
the cliff at t = 0, but in addition there is a downward (negative)
vertical velocity vy . The velocity vector at t = 0.50 s is
v = vxnd + vyne = 19.0 m>s2dn + 1-4.9 m>s2en
S
ure and heat
402
3.18 The initial velocity components v0x
Chapter 13 Gravitation
It’s usually simplest to take the initial position 1at t = 02 as the origin; then
and v0y of a projectile (such as a kicked
x0 = y0 = 0. This might be the position of a ball at the instant it leaves the hand
soccer ball)
are related to the initial speed
S
of
the person who throws it or the position of a bullet at the instant it leaves the
v0 and
a0 . star are
evaLuate: While the force magnitude F is tremendous, the
The components of the total
forceinitial
F onangle
the small
gun barrel.
magnitude of the resulting acceleration is not: a = F>m =
y
26
Fx = F1x + F2x = 1.81 * 10 N
3.17 shows
the
projectile
that starts at (or passes
2
11.87 Figure
* 1026 N2>11.00
* 1030
kg2trajectory
= 1.87 * of
10-4a m>s
. FurtherS
S
v0 = 4.72 * 1025 N
through)
theF
origin
time ttoward
= 0, the
along
withof its
position,
Fy = F1y + F2y
more,
the force
is not at
directed
center
mass
of the velocity, and velocity
S
components
large stars. at equal time intervals. The x-velocity vx is constant; the y-velocity vy
x two
582
ChapTer 17
Temperature
heat
The magnitude
of F
and its angleand
u (see
Fig. 13.5) are
inflUenced by The laTeST in edUcaTion reSearch
O
changes by equal amounts in equal times, just as if the projectile were launched
F = 2F x2 + F y2 = 211.81 * 1026 N22 + 14.72 * 1025 N22
vertically with the same initial y-velocity.
S
17.115 ... A hollow cylinder has length yL, inner radius a, and 17.116 You place 0.350 g of this cryoprotectant at 22°C
We can also represent the initial velocity v0 inbyconits magnitude v0 (the initial
26
= 1.87b, *
outer radius
and10the N
temperatures at the inner and outer surfaces tact with a cold plate that is maintained at the boiling temperature
speed) and its angle a0 with the positive x-axis (Fig. 3.18). In terms of these
S
are T2 and T1. (The
represent
an
insulated
hot-water
v
25
of
liquid
nitrogen
(77
K).
The
cryoprotectant
is
thermally
insuFy cylinder could
0
4.72 * 10 N
v0x and
of the in
initial
velocity are
the components
0y values
= arctan
arctan
= 14.6°
pipe.)u The
thermal =conductivity
of the26material
of which the cyl- lated quantities,
from everything
but the cold plate.
Use vthe
the table
F
* v10sin
N
v0y =
a
inder is made is k.x Derive an1.81
equation
for
(a)
the
total
heat
current
to
determine
how
much
heat
will
be
transferred
from
the
cryo0
0
through the walls of the cylinder; (b) the temperature variation protectant as it reaches thermal equilibrium with the cold plate.
a
v0x(c)=3.4v0*cos
v0y *= 10v05 sin
(3.18)
50
inside the cylinder walls. (c) Show that the equation0 for the total x (a) 1.5 * 105 J; (b) 2.9 * 105 J;
10a
J; (d) 4.4
J. a0
Data SpeakS SiDebarS, based on MasteringPhysics
heat current reduces to Eq. (17.21) for linear
v0x heat
= v0 flow
cos a0when 17.117 Careful measurements show that the specific heat of the
why Gravitational
forces
important
metadata,
alertphase
students
toare
the
statistically
common
the cylinder wall is very thin. (d) A steam pipe with a radius
of solid
depends
on
temperature
(Fig.most
P17.117).
How will the
Ifand
we
substitute
Eqs.
(3.18)
into to
Eqs.
(3.14)
through
(3.17) and set x0 = y0 = 0,
Examples
13.1
13.3
shows
that
gravitational
forces
negligible
2.00Gravitation
cm, carrying steam at 140°C, is surroundedComparing
by a cylindrical
actual
time
needed
for
this
cryoprotectant
come
to are
equilibrium
mistakes made in solving problems
on a given
topic.
we cold
get the
equations.
They
describe
theobjects
position and velocity of the
jacket with inner and outer radii 2.00 cm and 4.00
cm andordinary
made with
between
household-sized
objects
very
substantial
between
the
platefollowing
comparebut
with
the time
predicted
by using
the
When students were given a problem
projectile
ingravitation
Fig.
3.17that
atisany
time
of a type
of cork with thermal conductivity 4.00 *that
10-2
K. ofvalues
inIndeed,
the table?
Assume
all
than the
specific
areW>m
the #size
stars.
thevalues
mostt:other
important
force
on the
about superposition of gravitational
This forces,
in turnmore
is surrounded
by
a
cylindrical
jacket
made
of
a
brand stars,
heat (solid)
are correct.
The13.6).
actualIttime
(a) will be shorter;
(b) willour
scale
of
planets,
and galaxies
(Fig.
is responsible
for holding
than 60% gave an incorrect
-2
# K andand be
of Styrofoam
thermal
conductivity 2.70 * 10earth
W>m
(c)the
willplanets
be the in
same;
depends
on the
density
of the
response.with
Common
errors:
together
forlonger;
keeping
orbit(d)about
the sun.
The
mutual
gravhaving inner and outer radii 4.00 cm and 6.00 cmitational
(Fig. P17.115).
cryoprotectant.
attraction
between
different
parts
of
the
sun
compresses
material
at the
x = 1v0 cos a02t
(3.19)
● Assuming that equal-mass objects A and
The outer surface of the Styrofoam has a temperature of 15°C.
sun’s core to very high densities
and temperatures,
making it possible for nuclear
Coordinates
at time t of
B must exert equally strong gravitational
What is the temperature at a radius of 4.00 cm, where the two Figure P17.117
a projectile
(positive generate Speed
attraction on an object C (which is not
reactions oftoheat
take place there.
These reactions
the sun’sDirection
energy output,
Time
insulating
layers meet? (e) What
total rate of transfer Demo
Demois the
y-direction
at t = 0 at t = 0
# K)is upward,
true when A and B are different
distancesDemo which makes
c (J>kg
it possible for life
to exist
on earth and for you
to read these words.
out of afrom
2.00-m
length
of
pipe?
and
x
=
y
=
0
at
t
=
0)
C).
Pedagogy informed by daTa and reSearch
Data SpeakS
The gravitational force is5000
so important on the cosmic
because1 it 2acts
y = 1vscale
0 sin a02t - 2 gt
Neglecting to account for the vector
4000 contact between bodies. Electric
at a distance, without any direct
and magnetic
nature
of
force.
(To
add
two
forces
that
3000 property, but they are less important on astroFigure P17.115
forces have this same remarkable
point in different directions,
you
can’t
v
=
v
cos
a
Acceleration
PhET: Projectile Motion
x
0
0
2000
nomical scales because large
accumulations
just add the force magnitudes.)
Velocity
components atof matter are electrically neutral;
due to gravity:
6.00 that is, they contain equal amounts
1000
of
positive
and
negative
charge.
As
a
result,
the
Speed
Direction
time t of a projectile
Note g 7 0.
r =
4.00 cm
T
(°C)
0
at
t
=
0
at
t
=
0
(positive
y-direction
electric and magnetic forces
between
stars
planets
2.00 cm cm
-200 -150
-100or -50
0 are50very small or zero.
is
upward)
The strong and weak interactions that we discussed in Section 5.5 also actTime
at
Steam pipe
(3.20)
●
(3.21)
(3.22)
vy = v0 sin a0 - gt
a distance, but their
influence
is negligible
at distances
17.118
In another
experiment,
you placemuch
a layergreater
of thisthan
cryo-the
-14
m).
diameter of an atomic
nucleus
(aboutone
10 10 cm
.. cm
protectant
between
* 10
coldYou
plate
at motion of a large flywheel
9.88
DATA
aremaintained
analyzing the
roughly 1011 stars as well as gas, dust, and
A useful way to-describe
thatcold
actplate
atthat
a distance
is 0.800
in
of one
a field.
Onethe wheel starts from rest
40°C andforces
a second
ofhas
theradius
same
sizeterms
maintained
at liqm. In
test
run,
other matter. The entire assemblage is held
uid nitrogen’s
boiling
(77with
K).
Then
you
measure
the on
body sets up a disturbance
or field
at alltemperature
points
inturns
space,
and
the
force
thatacceleration.
acts
and
constant
angular
An
accelerometer
on
All
key
equationS
are
now
annotateD
to help
students
together by the mutual gravitational attracof heat transfer.
wants
tofirst
repeat
themeasures
experiment
a second body at arate
particular
point is Another
its response
toofthe
body’s
field atthe
that
thelab
rim
the
flywheel
magnitude of the resultant
tion of all the matter in the galaxy.
make
a
connection
between
a
conceptual
and
a
mathematical
PaSSagE ProbLEmS
but associated
uses cold plates
20 cm
20a cm,
one
atand
40°C
acceleration
of
point
on the
rim
of we
the flywheel as a function of
point. There is a field
withthat
eachareforce
that* acts
at aawith
distance,
so
understanding
of
physics.
andfields,
the other
at 77 K.
Howmagnetic
thick
does
the
of so
cryoprotectant
the
angle
ufields,
- ulayer
which
thewon’t
wheel has turned. You collect
0 through
refer
to
gravitational
electric
fields,
and
on.
We
BIO PrEsErving CElls aT Cold TEmPEraTurEs. In have to be so that the rate of heat transfer by conduction is the
these results:this chapter, so we won’t
76 the field concept for our study of gravitation
need
cryopreservation, M03_YOUN3610_14_SE_C03_067-100.indd
biological materials are cooled
to
a very low same as that when you use the smallerinplates?
(a) One-quarter the
discuss
further here. But in later chapters Uwe’ll
the field
is an2.50 3.00 3.50 4.00
0.50
1.00 concept
1.50 2.00
− U find
1rad 2that
temperature to slow down chemical reactions that
might itdamage
thickness; (b) half the thickness; 0(c) twice the thickness; (d) four
tool
for
describing
electric
and
magnetic
interactions.
the cells or tissues. It is important to prevent theextraordinarily
materials from powerful
2
times the thickness.
Cork
part of a spiral
13.6 Our solar system is
galaxy like this one, which
contains
Styrofoam
a 1m , s 2
0.678
1.07
1.52
1.98
2.45
2.92
3.39
Figure P9.90
9/11/14 10:45 AM
3.87
forming ice crystals during freezing. One method for preventing 17.119 To measure the specific heat in the liquid phase of a newly
4
formation
is to place
the material
protective
solution
called
aThe
graph
of new
a2 (in
m2>s
) versus 1u - u022 in (rad2).
test
your
of seCtioN
13.1
planet
Saturn
has
about
has length L, inner radius ice
a, and
17.116
You place
0.350 ginofa this
cryoprotectant
at uNDerstaNDiNG
22°C in
con- cryoprotectant,
developed
youConstruct
place
a sample
of the
cryoproteca
cryoprotectant.
Stated
values
of
the
thermal
properties
of
one
(a)
What
are
the
slope
and
y-intercept
of the straight
that
100
times
the
mass
of
the
earth
and
is
about
10
times
farther
from
the
sun
than
the
earth
ures at the inner and outer surfaces tact with a cold plate that is maintained at the boiling temperature
problemS
appear indrops
each chapter.
Theseline
datatant in contact with a cold Data
plate until
the solution’s temperature
cryoprotectant
are listed
here:(77 K). The cryoprotectant
gives
best
to the
data?
(b) the
Use
is. Compared
to the from
acceleration
of the earthtocaused
bythe
thepoint.
sun’sfit
gravitational
pull,
how the slope from part (a) to
d represent an insulated hot-water
of liquid
nitrogen
is thermally
insu-room temperature
its
freezing
Then
you
measure
based
reasoning
problems,
of which are context
find
the
angular
ofmany
the
is the
acceleration
of
Saturn due
the
sun’s
gravitation?
(i) acceleration
100
greater;
of the material of which the cyl- lated from everything but the cold plate.great
Use the
values
in theheat
table
transferred
to to
the
cold
plate.
If
the
system
isn’ttimes
sufficiently
iso-flywheel. (c) What is the lin1
1
ear(v)
speed
a point
rimexperimental
of the flywheelevidence,
when the wheel
rich,
require
students
tothe
use
- 20°C
(ii) 10 timesfrom
greater;
(iii)
the same;
(iv) 10
as great;
Melting point
❙ on
uation for (a) the total heat current
to determine how much
heat will be transferred
thelated
cryo100 asofgreat.
from
its room-temperature
surroundings,
what
will
be
the
effect
has turned
angle
of
135°? (d)
When the
flywheel has
er; (b) the temperature variation
equilibrium with the coldon
plate.
Latent heatprotectant
of fusion as it reaches
* 105 J>kg
2.80thermal
the measurement of thepresented
specific
heat?
(a)
The an
measured
specific
inthrough
a tabular
or graphical
format,
to formulate
5
5
5
through heat;
an angle
of measured
90.0°, what is the angle between the
ow that the equation for the total (a) 1.5 * 105 J; (b) 2.9
(c)# 3.4
J.will be greater than the turned
heat
actual specific
(b) the
4.5 ** 10
103 J;
J>kg
K * 10 J; (d) 4.4 * 10
conclusions.
velocity
of heat;
a point
itswill
rimbe
and the resultant acceleration
7.21) for linear heat flow when 17.117 Careful measurements show that
the specific
heatspecific
of the heat will be less than linear
the actual
specific
(c) on
there
13.2
weiGht
that point?of the cryoprotectant is so
d) A steam pipe with a radius of solid phase depends on temperature (Fig. P17.117). How will
the because the thermal of
no effect
conductivity
.. specific
DATA You
°C, is surrounded by a cylindrical actual time needed for this mcryoprotectant
come tothe
equilibrium
low; (d)ofthere
will be
effect9.89
on4.4
theas
heat,are
but rebuilding
the
tempera-a 1965 Chevrolet. To decide
Wetodefined
weight
a body
innoSection
the attractive
gravitational
whether
to replace the flywheel with a newer, lighter-weight one,
i 2.00 cm and 4.00 cm and made with the cold plate compare with the time
predicted
by
using
the
ture
of the
freezing
point
will
change.
force
exerted
on
it
by
the
earth.
We
can
now
broaden
our
definition
and
say that
you want to determine the moment of inertia of the original,
onductivity 4.00 * 10-2 W>m # K. values in the table? Assume that all values
than
specific
the other
weight
of the
a body
is the total gravitational
force exertedflywheel.
on the body
by aall
35.6-cm-diameter
It
is
not
uniform disk, so you can’t
cylindrical jacket made of a brand heat (solid) are correct. The actual time (a) will be shorter; (b) will
1
2
other bodies in the universe. When the body
is=near
the
surface
of
the
earth,of inertia. You remove the
-2
use
I
MR
to
calculate
the
moment
#
2
uctivity 2.70 * 10 W>m K and be longer; (c) will be the same; (d) depends on the density of the
we can ignore all other gravitational forcesflywheel
and consider
thecar
weight
as low-friction
just the
from the
and use
bearings to mount it
0 cm and 6.00 cm (Fig. P17.115). cryoprotectant.
earth’s gravitational attraction. At the surface
the moon stationary
we consider
body’s
on of
a horizontal,
rod athat
passes through the center of
foam has a temperature of 15°C.
weight to be the gravitational attraction of the
andwhich
so on.can then rotate freely (about 2 m above the
PhET: Lunar Lander
themoon,
flywheel,
adius of 4.00 cm, where the two
Figure P17.117
ground). After gluing one end of a long piece of flexible fishing
is the total rate of transfer of heat
c (J>kg # K)
line to the rim of the flywheel, you wrap the line a number of
5000
turns around the rim and suspend a 5.60-kg metal block from
4000
the free end of the line. When you release the block from rest,
3000
it descends as the flywheel rotates. With high-speed photography
2000
you measure the distance d the block has moved downward as a
6.00
function of the time since it was released. The
equation for the
1000 402
M13_YOUN3610_14_SE_C13_398-432.indd
30/06/14 12:46 PM
=
4.00 cm
graph shown in Fig. P9.89 that gives a good fit to the data points
T
(°C)
0
.00 cm cm
-200 -150 -100 -50
0
50
is d = 1165 cm>s22t 2. (a) Based on the graph, does the block fall
am pipe
with constant acceleration? Explain. (b) Use the graph to calculate
17.118 In another experiment, you place a layer of this cryothe speed of the block when it has descended 1.50 m. (c) Apply
Each
chapter
includes
three
to
five
paSSage
problemS,
am
protectant between one 10 cm * 10 cm cold plate maintained at
conservation of mechanical energy to the system of flywheel and
whichand
follow
thecold
format
thesize
MCATs.
These
problems
-40°C
a second
plateused
of the in
same
maintained
at liqblock to calculate the moment of inertia of the flywheel. (d) You
uid
nitrogen’s
boilingto
temperature
(77 multiple
K). Then you
measure
are relieved that the fishing line doesn’t break.
Apply
M17_YOUN3610_14_SE_C17_545-583.indd
582
7/30/14 12:39
PMNewton’s
require
students
investigate
aspects
of the
a real-life
rate of heat transfer. Another lab wants to repeat the experiment
second law to the block to find the tension in the line as the block
physical
situation,
typically
biological
in
nature,
as
described
but uses cold plates that are 20 cm * 20 cm, with one at - 40°C
descended.
in athe
reading
and
other atpassage.
77 K. How thick does the layer of cryoprotectant
T Cold TEmPEraTurEs. In have to be so that the rate of heat transfer by conduction is the
Figure P9.89
terials are cooled to a very low same as that when you use the smaller plates? (a) One-quarter the
d (cm)
mical reactions that might damage thickness; (b) half the thickness; (c) twice the thickness; (d) four
200
ant to prevent the materials from times the thickness.
... CALC
tern of tiny pits a
rim of the disc. A
scanned at a con
radius of the trac
of the disc must
Let’s see what an
The equation of a
of the spiral at u
radius of the spir
CD to be positive
turns and u incre
angle du, the dist
the above expres
tance s scanned
through which th
at a constant linea
to vt. Use this to
solutions for u ; c
the solution to ch
angular velocity
of time. Is az con
is 25.0 mm, the t
and the playing t
of revolutions ma
from parts (c) and
(in rad>s2) versus
9.91
paSSage p
BIO The Spinn
are freshwater fi
uniform cylinder
pensates for its s
mouth and then
PerSonALize LeArning with MASteringPhySiCS
M
asteringPhysics® from Pearson is the leading online homework, tutorial, and assessment system,
designed to improve results by engaging students before, during, and after class with powerful
content. Instructors can now ensure that students arrive ready to learn by assigning educationally
effective content before class, and encourage critical thinking and retention with in-class resources such
as Learning Catalytics. Students can further master concepts after class through traditional and adaptive
homework assignments that provide hints and answer-specific feedback. The Mastering gradebook records
scores for all automatically graded assignments in one place, while diagnostic tools give instructors access
to rich data to assess student understanding and misconceptions.
Mastering brings learning full circle by continuously adapting to each student and making learning more
personal than ever—before, during, and after class.
Before CLASS
interaCtive PreLeCture videos address
the rapidly growing movement
toward pre-lecture teaching
and flipped classrooms.
These videos provide a
conceptual introduction
to key topics. Embedded
assessment helps students to
prepare before lecture and
instructors to identify student
misconceptions.
During CLASS
Learning CataLytiCs™ is a “bring your own
device” student engagement, assessment, and
classroom intelligence system. With Learning
Catalytics you can:
•
•
•
•
•
•
Assess students in real time, using open-ended
tasks to probe student understanding.
Understand immediately where students are
and adjust your lecture accordingly.
Improve your students’ critical-thinking skills.
Access rich analytics to understand student
performance.
Add your own questions to make Learning
Catalytics fit your course exactly.
Manage student interactions with intelligent
grouping and timing.
Pre-LeCture ConCePt
Questions check familiarity with
key concepts, prompting students
to do their assigned reading prior
to coming to class. These quizzes
keep students on track, keep them
more engaged in lecture, and help
you spot the concepts with which
they have the most difficulty. Openended essay questions help students
identify what they find most difficult
about a concept, better informing
you and assisting with “just-in-time”
teaching.
beFoRe, DuRing, AnD AFteR CLASS
AFteR CLASS
tUtorials featuring specific wronganswer feedback, hints, and a wide variety
of educationally effective content guide
your students through the toughest topics
in physics. The hallmark Hints and Feedback
offer instruction similar to what students
would experience in an office hour, allowing
them to learn from their mistakes without
being given the answer.
adaptive follow-Ups are personalized assignments
that pair Mastering’s powerful content with Knewton’s adaptive
learning engine to provide personalized help to students.
These assignments address common student misconceptions
and topics students struggled with on assigned homework,
including core prerequisite topics.
video tUtor demonstrations, available in
the Study Area and in the Item Library and accessible
by QR code in the textbook, feature “pause-andpredict” demonstrations of key physics concepts
as assessment to engage students actively in
understanding key concepts. New VTDs build on
the existing collection, adding new topics for a more
robust set of demonstrations.
video tUtor solUtions are tied to each worked example and Bridging
Problem in the textbook and can be accessed through MasteringPhysics or from
QR codes in the textbook. They walk students through the problem-solving
process, providing a virtual teaching assistant on a round-the-clock basis.
About tHe AutHoRS
Roger A. Freedman is a Lecturer in Physics at the University of California, Santa
Barbara. He was an undergraduate at the University of California campuses in San
Diego and Los Angeles and did his doctoral research in nuclear theory at Stanford
University under the direction of Professor J. Dirk Walecka. Dr. Freedman came to
UCSB in 1981 after three years of teaching and doing research at the University of
Washington.
At UCSB, Dr. Freedman has taught in both the Department of Physics and the
College of Creative Studies, a branch of the university intended for highly gifted and
motivated undergraduates. He has published research in nuclear physics, elementary
particle physics, and laser physics. In recent years, he has worked to make physics
lectures a more interactive experience through the use of classroom response systems
and pre-lecture videos.
In the 1970s Dr. Freedman worked as a comic book letterer and helped organize the
San Diego Comic-Con (now the world’s largest popular culture convention) during
its first few years. Today, when not in the classroom or slaving over a computer,
Dr. Freedman can be found either flying (he holds a commercial pilot’s license) or
with his wife, Caroline, cheering on the rowers of UCSB Men’s and Women’s Crew.
in MeMOriAM: hugh yOung (1930–2013)
dumperina
Hugh D. Young was Emeritus Professor of Physics at Carnegie Mellon University. He
earned both his undergraduate and graduate degrees from that university. He earned his
Ph.D. in fundamental particle theory under the direction of the late Richard Cutkosky.
Dr. Young joined the faculty of Carnegie Mellon in 1956 and retired in 2004. He also
had two visiting professorships at the University of California, Berkeley.
Dr. Young’s career was centered entirely on undergraduate education. He wrote
several undergraduate-level textbooks, and in 1973 he became a coauthor with Francis
Sears and Mark Zemansky for their well-known introductory textbooks. In addition
to his role on Sears and Zemansky’s University Physics, he was the author of Sears
and Zemansky’s College Physics.
Dr. Young earned a bachelor’s degree in organ performance from Carnegie Mellon
in 1972 and spent several years as Associate Organist at St. Paul’s Cathedral in
Pittsburgh. He often ventured into the wilderness to hike, climb, or go caving with
students in Carnegie Mellon’s Explorers Club, which he founded as a graduate student
and later advised. Dr. Young and his wife, Alice, hosted up to 50 students each year
for Thanksgiving dinners in their home.
Always gracious, Dr. Young expressed his appreciation earnestly: “I want to extend
my heartfelt thanks to my colleagues at Carnegie Mellon, especially Professors Robert
Kraemer, Bruce Sherwood, Ruth Chabay, Helmut Vogel, and Brian Quinn, for many
stimulating discussions about physics pedagogy and for their support and encouragement during the writing of several successive editions of this book. I am equally
indebted to the many generations of Carnegie Mellon students who have helped me
learn what good teaching and good writing are, by showing me what works and what
doesn’t. It is always a joy and a privilege to express my gratitude to my wife, Alice,
and our children, Gretchen and Rebecca, for their love, support, and emotional sustenance during the writing of several successive editions of this book. May all men and
women be blessed with love such as theirs.” We at Pearson appreciated his professionalism, good nature, and collaboration. He will be missed.
A. Lewis Ford is Professor of Physics at Texas A&M University. He received a B.A.
from Rice University in 1968 and a Ph.D. in chemical physics from the University of
Texas at Austin in 1972. After a one-year postdoc at Harvard University, he joined the
Texas A&M physics faculty in 1973 and has been there ever since. Professor Ford has
specialized in theoretical atomic physics—in particular, atomic collisions. At Texas
A&M he has taught a variety of undergraduate and graduate courses, but primarily
introductory physics.
tO the student
HoW to SuCCeeD in PHYSiCS
bY ReALLY tRYing
Mark Hollabaugh, Normandale Community College, Emeritus
Physics encompasses the large and the small, the old and the new. From the atom
to galaxies, from electrical circuitry to aerodynamics, physics is very much a
part of the world around us. You probably are taking this introductory course in
calculus-based physics because it is required for subsequent courses that you plan
to take in preparation for a career in science or engineering. Your professor wants
you to learn physics and to enjoy the experience. He or she is very interested in
helping you learn this fascinating subject. That is part of the reason your professor chose this textbook for your course. That is also the reason Drs. Young and
Freedman asked me to write this introductory section. We want you to succeed!
The purpose of this section of University Physics is to give you some ideas
that will assist your learning. Specific suggestions on how to use the textbook
will follow a brief discussion of general study habits and strategies.
prepArAtiOn fOr this cOurse
If you had high school physics, you will probably learn concepts faster than those
who have not because you will be familiar with the language of physics. If English
is a second language for you, keep a glossary of new terms that you encounter
and make sure you understand how they are used in physics. Likewise, if you are
further along in your mathematics courses, you will pick up the mathematical
aspects of physics faster. Even if your mathematics is adequate, you may find
a book such as Arnold D. Pickar’s Preparing for General Physics: Math Skill
Drills and Other Useful Help (Calculus Version) to be useful. Your professor may
assign sections of this math review to assist your learning.
leArning tO leArn
Each of us has a different learning style and a preferred means of learning.
Understanding your own learning style will help you to focus on aspects of
physics that may give you difficulty and to use those components of your course
that will help you overcome the difficulty. Obviously you will want to spend
more time on those aspects that give you the most trouble. If you learn by hearing,
lectures will be very important. If you learn by explaining, then working with
other students will be useful to you. If solving problems is difficult for you, spend
more time learning how to solve problems. Also, it is important to understand
and develop good study habits. Perhaps the most important thing you can do for
yourself is set aside adequate, regularly scheduled study time in a distraction-free
environment.
Answer the following questions for yourself:
• Am I able to use fundamental mathematical concepts from algebra, geometry,
and trigonometry? (If not, plan a program of review with help from your
professor.)
• In similar courses, what activity has given me the most trouble? (Spend more
time on this.) What has been the easiest for me? (Do this first; it will build
your confidence.)
• Do I understand the material better if I read the book before or after the
lecture? (You may learn best by skimming the material, going to lecture, and
then undertaking an in-depth reading.)
ix
x
How to Succeed in PHySicS by really trying
• Do I spend adequate time studying physics? (A rule of thumb for a class like
this is to devote, on average, 2.5 hours out of class for each hour in class. For
a course that meets 5 hours each week, that means you should spend about
10 to 15 hours per week studying physics.)
• Do I study physics every day? (Spread that 10 to 15 hours out over an entire
week!) At what time of the day am I at my best for studying physics? (Pick a
specific time of the day and stick to it.)
• Do I work in a quiet place where I can maintain my focus? (Distractions will
break your routine and cause you to miss important points.)
WOrking With Others
Scientists or engineers seldom work in isolation from one another but rather work
cooperatively. You will learn more physics and have more fun doing it if you
work with other students. Some professors may formalize the use of cooperative learning or facilitate the formation of study groups. You may wish to form
your own informal study group with members of your class. Use e-mail to keep
in touch with one another. Your study group is an excellent resource when you
review for exams.
lectures And tAking nOtes
An important component of any college course is the lecture. In physics this is
especially important, because your professor will frequently do demonstrations
of physical principles, run computer simulations, or show video clips. All of
these are learning activities that will help you understand the basic principles of
physics. Don’t miss lectures. If for some reason you do, ask a friend or member
of your study group to provide you with notes and let you know what happened.
Take your class notes in outline form, and fill in the details later. It can be very
difficult to take word-for-word notes, so just write down key ideas. Your professor may use a diagram from the textbook. Leave a space in your notes and add
the diagram later. After class, edit your notes, filling in any gaps or omissions and
noting things that you need to study further. Make references to the textbook by
page, equation number, or section number.
Ask questions in class, or see your professor during office hours. Remember
that the only “dumb” question is the one that is not asked. Your college may have
teaching assistants or peer tutors who are available to help you with any difficulties.
exAMinAtiOns
Taking an examination is stressful. But if you feel adequately prepared and are
well rested, your stress will be lessened. Preparing for an exam is a continuous
process; it begins the moment the previous exam is over. You should immediately
go over the exam to understand any mistakes you made. If you worked a problem
and made substantial errors, try this: Take a piece of paper and divide it down the
middle with a line from top to bottom. In one column, write the proper solution to
the problem. In the other column, write what you did and why, if you know, and
why your solution was incorrect. If you are uncertain why you made your mistake or how to avoid making it again, talk with your professor. Physics constantly
builds on fundamental ideas, and it is important to correct any misunderstandings immediately. Warning: Although cramming at the last minute may get you
through the present exam, you will not adequately retain the concepts for use on
the next exam.
tO the instructOr
PReFACe
This book is the product of six and a half decades of leadership and innovation
in physics education. When the first edition of University Physics by Francis
W. Sears and Mark W. Zemansky was published in 1949, it was revolutionary
among calculus-based physics textbooks in its emphasis on the fundamental
principles of physics and how to apply them. The success of University Physics
with generations of several million students and educators around the world is a
testament to the merits of this approach and to the many innovations it has introduced subsequently.
In preparing this new Fourteenth Edition, we have further augmented and
developed University Physics to assimilate the best ideas from education research
with enhanced problem-solving instruction, pioneering visual and conceptual
pedagogy, all-new categories of end-of-chapter problems, and the most pedagogically proven and widely used online homework and tutorial system in the
world.
neW tO this editiOn
• All key equations now include annotations that describe the equation and
explain the meanings of the symbols in the equation. These annotations help
promote in-depth processing of information and greater recall.
• DATA SPEAKS sidebars in each chapter, based on data captured from thousands of students, alert students to the statistically most common mistakes
students make when working problems on related topics in MasteringPhysics.
• Updated modern physics content includes sections on quantum measurement (Chapter 40) and quantum entanglement (Chapter 41), as well as recent
data on the Higgs boson and cosmic background radiation (Chapter 44).
• Additional bioscience applications appear throughout the text, mostly in the
form of marginal photos with explanatory captions, to help students see how
physics is connected to many breakthroughs and discoveries in the biosciences.
• The text has been streamlined with tighter and more focused language.
• Based on data from MasteringPhysics, changes to the end-of-chapter content
include the following:
• 25%–30% of problems are new or revised.
• Most chapters include six to ten biosciences-related problems.
• The number of context-rich problems is increased to facilitate the greater
learning gains that they can offer.
• Three new DATA problems appear in each chapter. These typically contextrich, data-based reasoning problems require students to use experimental
evidence, presented in a tabular or graphical format, to formulate conclusions.
• Each chapter now includes three to five new Passage Problems, which
follow the format that is used in the MCATs. These problems require students
to investigate multiple aspects of a real-life physical situation, typically
biological in nature, that is described in a reading passage.
• Looking back at ... essential past concepts are listed at the beginning of each
chapter, so that students know what they need to have mastered before digging
into the current chapter.
Standard, Extended,
and Three-Volume Editions
With MasteringPhysics:
• Standard Edition: chapters 1–37
(iSbn 978-0-13-409650-6)
• Extended Edition: chapters 1–44
(iSbn 978-0-321-98258-2)
Without MasteringPhysics:
• Standard Edition: chapters 1–37
(iSbn 978-0-13-396929-0)
• Extended Edition: chapters 1–44
(iSbn 978-0-321-97361-0)
• Volume 1: chapters 1–20
(iSbn 978-0-13-397804-9)
• Volume 2: chapters 21–37
(iSbn 978-0-13-397800-1)
• Volume 3: chapters 37–44
(iSbn 978-0-13-397802-5)
xi
xii
PreFace
key feAtures Of University Physics
Demo
• More than 620 QR codes throughout the book allow students to use a mobile
phone to watch an interactive video of a physics instructor giving a relevant
physics demonstration (Video Tutor Demonstration) or showing a narrated
and animated worked Example (Video Tutor Solution).
All of these videos also play directly through links within the Pearson eText
as well as the Study Area within MasteringPhysics.
• End-of-chapter Bridging Problems, many revised, provide a transition between the single-concept Examples and the more challenging end-of-chapter
problems. Each Bridging Problem poses a difficult, multiconcept problem
that typically incorporates physics from earlier chapters. A skeleton Solution
Guide, consisting of questions and hints, helps train students to approach and
solve challenging problems with confidence.
• Deep and extensive problem sets cover a wide range of difficulty (with blue
dots to indicate relative difficulty level) and exercise both physical understanding and problem-solving expertise. Many problems are based on complex
real-life situations.
• This textbook offers more Examples and Conceptual Examples than most
other leading calculus-based textbooks, allowing students to explore problemsolving challenges that are not addressed in other textbooks.
• A research-based problem-solving approach (Identify, Set Up, Execute,
Evaluate) is used in every Example as well as in the Problem-Solving
Strategies, in the Bridging Problems, and throughout the Instructor’s Solutions
Manual and the Study Guide. This consistent approach teaches students to
tackle problems thoughtfully rather than cutting straight to the math.
• Problem-Solving Strategies coach students in how to approach specific types
of problems.
• The figures use a simplified graphical style to focus on the physics of a situation, and they incorporate more explanatory annotations than in the previous
edition. Both techniques have been demonstrated to have a strong positive
effect on learning.
• Many figures that illustrate Example solutions take the form of black-and-white
pencil sketches, which directly represent what a student should draw in solving
such problems themselves.
• The popular Caution paragraphs focus on typical misconceptions and student
problem areas.
• End-of-section Test Your Understanding questions let students check their
grasp of the material and use a multiple-choice or ranking-task format to
probe for common misconceptions.
• Visual Summaries at the end of each chapter present the key ideas in words,
equations, and thumbnail pictures, helping students review more effectively.
• Approximately 70 PhET simulations are linked to the Pearson eText and
provided in the Study Area of the MasteringPhysics website (with icons in the
printed book). These powerful simulations allow students to interact productively with the physics concepts they are learning. PhET clicker questions are
also included on the Instructor’s Resource DVD.
instructOr’s suppleMents
Note: For convenience, all of the following instructor’s supplements (except
for the Instructor’s Resource DVD) can be downloaded from the Instructor
Resources Area accessed via MasteringPhysics (www.masteringphysics.com).
The Instructor’s Solutions Manual, prepared by A. Lewis Ford (Texas A&M
University) and Wayne Anderson, contains complete and detailed solutions to all
end-of-chapter problems. All solutions follow consistently the same Identify/Set
Up/Execute/Evaluate problem-solving framework used in the textbook. Download
PREFACE
only from the MasteringPhysics Instructor Area or from the Instructor Resource
Center (www.pearsonhighered.com/irc).
The cross-platform Instructor’s Resource DVD (978-0-13-398364-7) provides
a comprehensive library of approximately 350 applets from ActivPhysics OnLine
as well as all art and photos from the textbook in JPEG and PowerPoint formats.
In addition, all of the key equations, problem-solving strategies, tables, and chapter summaries are provided in JPEGs and editable Word format, and all of the
new Data Speaks boxes are offered in JPEGs. In-class weekly multiple-choice
questions for use with various Classroom Response Systems (CRS) are also
provided, based on the Test Your Understanding questions and chapter-opening
questions in the text. Written by Roger Freedman, many new CRS questions that
increase in difficulty level have been added. Lecture outlines and PhET clicker
questions, both in PowerPoint format, are also included along with about 70 PhET
simulations and the Video Tutor Demonstrations (interactive video demonstrations) that are linked to QR codes throughout the textbook.
MasteringPhysics® (www.masteringphysics.com) from Pearson is the leading
online teaching and learning system designed to improve results by engaging
students before, during, and after class with powerful content. Ensure that students arrive ready to learn by assigning educationally effective content before
class, and encourage critical thinking and retention with in-class resources such
as Learning Catalytics. Students can further master concepts after class through
traditional homework assignments that provide hints and answer-specific feedback. The Mastering gradebook records scores for all automatically graded assignments, while diagnostic tools give instructors access to rich data to assess
student understanding and misconceptions.
Mastering brings learning full circle by continuously adapting to each student
and making learning more personal than ever—before, during, and after class.
• NEW! The Mastering Instructor Resources Area contains all of the contents of the Instructor’s Resource DVD—lecture outlines; Classroom Response
System questions; images, tables, key equations, problem-solving strategies,
Data Speaks boxes, and chapter summaries from the textbook; access to the
Instructor’s Solutions Manual, Test Bank, ActivPhysics Online—and much
more.
• NEW! Pre-lecture Videos are assignable interactive videos that introduce
students to key topics before they come to class. Each one includes assessment
that feeds to the gradebook and alerts the instructor to potential trouble spots
for students.
• Pre-lecture Concept Questions check students’ familiarity with key concepts,
prompting students to do their assigned reading before they come to class.
These quizzes keep students on track, keep them more engaged in lecture, and
help you spot the concepts that students find the most difficult.
• NEW! Learning Catalytics is a “bring your own device” student engagement,
assessment, and classroom intelligence system that allows you to assess students in real time, understand immediately where they are and adjust your
lecture accordingly, improve their critical-thinking skills, access rich analytics
to understand student performance, add your own questions to fit your course
exactly, and manage student interactions with intelligent grouping and timing.
Learning Catalytics can be used both during and after class.
• NEW! Adaptive Follow-Ups allow Mastering to adapt continuously to each
student, making learning more personal than ever. These assignments pair
Mastering’s powerful content with Knewton’s adaptive learning engine to
provide personalized help to students before misconceptions take hold. They
are based on each student’s performance on homework assignments and on all
work in the course to date, including core prerequisite topics.
xiii
xiv
PreFace
• Video Tutor Demonstrations, linked to QR codes in the textbook, feature
“Pause and predict” videos of key physics concepts that ask students to submit
a prediction before they see the outcome. These interactive videos are available in the Study Area of Mastering and in the Pearson eText.
• Video Tutor Solutions are linked to QR codes in the textbook. In these videos,
which are available in the Study Area of Mastering and in the Pearson eText,
an instructor explains and solves each worked example and Bridging Problem.
• NEW! An Alternative Problem Set in the Item Library of Mastering includes
hundreds of new end-of-chapter questions and problems to offer instructors a
wealth of options.
• NEW! Physics/Biology Tutorials for MasteringPhysics are assignable,
multipart tutorials that emphasize biological processes and structures but also
teach the physics principles that underlie them. They contain assessment questions that are based on the core competencies outlined in the 2015 MCAT.
• PhET Simulations (from the PhET project at the University of Colorado) are
interactive, research-based simulations of physical phenomena. These tutorials,
correlated to specific topics in the textbook, are available in the Pearson eText
and in the Study Area within www.masteringphysics.com.
• ActivPhysics OnLine™ (which is accessed through the Study Area and
Instructor Resources within www.masteringphysics.com) provides a comprehensive library of approximately 350 tried and tested ActivPhysics applets
updated for web delivery.
• Mastering’s powerful gradebook records all scores for automatically graded
assignments. Struggling students and challenging assignments are highlighted
in red, giving you an at-a-glance view of potential hurdles in the course. With
a single click, charts summarize the most difficult problems, identify vulnerable
students, and show the grade distribution, allowing for just-in-time teaching to
address student misconceptions.
• Learning Management System (LMS) Integration gives seamless access
to modified Mastering. Having all of your course materials and communications in one place makes life less complicated for you and your students.
We’ve made it easier to link from within your LMS to modified Mastering
and provide solutions, regardless of your LMS platform. With seamless, single
sign-on your students will gain access to the personalized learning resources
that make studying more efficient and more effective. You can access modified
Mastering assignments, rosters, and resources and synchronize grades from
modified Mastering with LMS.
• The Test Bank contains more than 2000 high-quality problems, with a range
of multiple-choice, true/false, short-answer, and regular homework-type questions. Test files are provided both in TestGen (an easy-to-use, fully networkable program for creating and editing quizzes and exams) and in Word format.
Download only from the MasteringPhysics Instructor Resources Area or
from the Instructor Resources Center (www.pearsonhighered.com/irc).
MasteringPhysics enables instructors to:
• Quickly build homework assignments that combine regular end-of-chapter
problems and tutoring (through additional multistep tutorial problems that
offer wrong-answer feedback and simpler problems upon request).
• Expand homework to include the widest range of automatically graded activities available—from numerical problems with randomized values, through
algebraic answers, to free-hand drawing.
• Choose from a wide range of nationally pre-tested problems that provide
accurate estimates of time to complete and difficulty.
• After an assignment is completed, quickly identify not only the problems
that were the trickiest for students but also the individual problem types with
which students had trouble.
PREFACE
• Compare class results against the system’s worldwide average for each problem assigned, to identify issues to be addressed with just-in-time teaching.
• Check the work of an individual student in detail, including the time spent on
each problem, what wrong answers were submitted at each step, how much
help was asked for, and how many practice problems were worked.
Student’S SupplementS
The Student’s Study Guide by Laird Kramer reinforces the textbook’s emphasis
on problem-solving strategies and student misconceptions. The Study Guide for
Volume 1 (978-0-13-398361-6) covers Chapters 1–20, and the Study Guide for
Volumes 2 and 3 (978-0-13-398360-9) covers Chapters 21–44.
The Student’s Solutions Manual by A. Lewis Ford (Texas A&M University)
and Wayne Anderson contains detailed, step-by-step solutions to more than half
of the odd-numbered end-of-chapter problems from the textbook. All solutions
follow consistently the same Identify/Set Up/Execute/Evaluate problem-solving
framework used in the textbook. The Student’s Solutions Manual for Volume 1
(978-0-13-398171-1) covers Chapters 1–20, and the Student’s Solutions Manual
for Volumes 2 and 3 (978-0-13-396928-3) covers Chapters 21–44.
MasteringPhysics® (www.masteringphysics.com) is a homework, tutorial, and
assessment system based on years of research into how students work physics
problems and precisely where they need help. Studies show that students who
use MasteringPhysics compared to handwritten homework significantly increase
their scores. MasteringPhysics achieves this improvement by providing students
with instantaneous feedback specific to their wrong answers, simpler sub-problems
upon request when they get stuck, and partial credit for their method(s). This individualized, 24/7 Socratic tutoring is recommended by nine out of ten students to
their peers as the most effective and time-efficient way to study.
Pearson eText is available through MasteringPhysics either automatically, when
MasteringPhysics is packaged with new books, or as a purchased upgrade online.
Allowing students access to the text wherever they have access to the Internet,
Pearson eText comprises the full text, including figures that can be enlarged for
better viewing. With eText, students are also able to pop up definitions and terms
to help with vocabulary and the reading of the material. Students can also take
notes in eText by using the annotation feature at the top of each page.
Pearson Tutor Services (www.pearsontutorservices.com). Each student’s subscription to MasteringPhysics also contains complimentary access to Pearson Tutor
Services, powered by Smarthinking, Inc. By logging in with their MasteringPhysics
ID and password, students are connected to highly qualified e-instructors who
provide additional interactive online tutoring on the major concepts of physics.
Some restrictions apply; the offer is subject to change.
TIPERs (Tasks Inspired by Physics Education Research) are workbooks that
give students the practice they need to develop reasoning about physics and that
promote a conceptual understanding of problem solving:
• NEW! TIPERs: Sensemaking Tasks for Introductory Physics (978-0-13-285458-0)
by Curtis Hieggelke, Stephen Kanim, David Maloney, and Thomas O’Kuma
• Newtonian Tasks Inspired by Physics Education Research: nTIPERs (978-0-32175375-5) by Curtis Hieggelke, David Maloney, and Stephen Kanim
• E&M TIPERs: Electricity & Magnetism Tasks (978-0-13-185499-4) by Curtis
Hieggelke, David Maloney, Thomas O’Kuma, and Stephen Kanim
Tutorials in Introductory Physics (978-0-13-097069-5) by Lillian C. McDermott
and Peter S. Schaffer presents a series of physics tutorials designed by a leading
physics education research group. Emphasizing the development of concepts and
scientific reasoning skills, the tutorials focus on the specific conceptual and reasoning difficulties that students tend to encounter.
xv
xvi
PreFace
AcknOWledgMents
I would like to thank the hundreds of reviewers and colleagues who have offered
valuable comments and suggestions over the life of this textbook. The continuing
success of University Physics is due in large measure to their contributions.
Miah Adel (U. of Arkansas at Pine Bluff), Edward Adelson (Ohio State U.), Julie
Alexander (Camosun C.), Ralph Alexander (U. of Missouri at Rolla), J. G. Anderson,
R. S. Anderson, Wayne Anderson (Sacramento City C.), Sanjeev Arora (Fort Valley State U.),
Alex Azima (Lansing Comm. C.), Dilip Balamore (Nassau Comm. C.), Harold Bale
(U. of North Dakota), Arun Bansil (Northeastern U.), John Barach (Vanderbilt U.), J. D.
Barnett, H. H. Barschall, Albert Bartlett (U. of Colorado), Marshall Bartlett (Hollins U.),
Paul Baum (CUNY, Queens C.), Frederick Becchetti (U. of Michigan), B. Bederson,
David Bennum (U. of Nevada, Reno), Lev I. Berger (San Diego State U.), Angela Biselli
(Fairfield U.), Robert Boeke (William Rainey Harper C.), Bram Boroson (Clayton State U.),
S. Borowitz, A. C. Braden, James Brooks (Boston U.), Nicholas E. Brown (California
Polytechnic State U., San Luis Obispo), Tony Buffa (California Polytechnic State U.,
San Luis Obispo), Shane Burns (Colorado C.), A. Capecelatro, Michael Cardamone
(Pennsylvania State U.), Duane Carmony (Purdue U.), Troy Carter (UCLA), P. Catranides,
John Cerne (SUNY at Buffalo), Shinil Cho (La Roche C.), Tim Chupp (U. of Michigan),
Roger Clapp (U. of South Florida), William M. Cloud (Eastern Illinois U.), Leonard
Cohen (Drexel U.), W. R. Coker (U. of Texas, Austin), Malcolm D. Cole (U. of Missouri
at Rolla), H. Conrad, David Cook (Lawrence U.), Gayl Cook (U. of Colorado), Hans
Courant (U. of Minnesota), Carl Covatto (Arizona State U.), Bruce A. Craver (U. of
Dayton), Larry Curtis (U. of Toledo), Jai Dahiya (Southeast Missouri State U.), Dedra
Demaree (Georgetown U.), Steve Detweiler (U. of Florida), George Dixon (Oklahoma
State U.), Steve Drasco (Grinnell C.), Donald S. Duncan, Boyd Edwards (West Virginia U.),
Robert Eisenstein (Carnegie Mellon U.), Amy Emerson Missourn (Virginia Institute of
Technology), Olena Erhardt (Richland C.), William Faissler (Northeastern U.), Gregory
Falabella (Wagner C.), William Fasnacht (U.S. Naval Academy), Paul Feldker (St. Louis
Comm. C.), Carlos Figueroa (Cabrillo C.), L. H. Fisher, Neil Fletcher (Florida State U.),
Allen Flora (Hood C.), Robert Folk, Peter Fong (Emory U.), A. Lewis Ford (Texas A&M U.),
D. Frantszog, James R. Gaines (Ohio State U.), Solomon Gartenhaus (Purdue U.), Ron
Gautreau (New Jersey Institute of Technology), J. David Gavenda (U. of Texas, Austin),
Dennis Gay (U. of North Florida), Elizabeth George (Wittenberg U.), James Gerhart
(U. of Washington), N. S. Gingrich, J. L. Glathart, S. Goodwin, Rich Gottfried (Frederick
Comm. C.), Walter S. Gray (U. of Michigan), Paul Gresser (U. of Maryland), Benjamin
Grinstein (UC, San Diego), Howard Grotch (Pennsylvania State U.), John Gruber (San
Jose State U.), Graham D. Gutsche (U.S. Naval Academy), Michael J. Harrison (Michigan
State U.), Harold Hart (Western Illinois U.), Howard Hayden (U. of Connecticut), Carl
Helrich (Goshen C.), Andrew Hirsch (Purdue U.), Linda Hirst (UC, Merced), Laurent
Hodges (Iowa State U.), C. D. Hodgman, Elizabeth Holden (U. of Wisconsin, Platteville),
Michael Hones (Villanova U.), Keith Honey (West Virginia Institute of Technology),
Gregory Hood (Tidewater Comm. C.), John Hubisz (North Carolina State U.), Eric Hudson
(Pennsylvania State U.), M. Iona, Bob Jacobsen (UC, Berkeley), John Jaszczak (Michigan
Technical U.), Alvin Jenkins (North Carolina State U.), Charles Johnson (South Georgia
State C.), Robert P. Johnson (UC, Santa Cruz), Lorella Jones (U. of Illinois), Manoj
Kaplinghat (UC, Irvine), John Karchek (GMI Engineering & Management Institute),
Thomas Keil (Worcester Polytechnic Institute), Robert Kraemer (Carnegie Mellon U.),
Jean P. Krisch (U. of Michigan), Robert A. Kromhout, Andrew Kunz (Marquette U.),
Charles Lane (Berry C.), Stewart Langton (U. of Victoria), Thomas N. Lawrence (Texas
State U.), Robert J. Lee, Alfred Leitner (Rensselaer Polytechnic U.), Frederic Liebrand
(Walla Walla U.), Gerald P. Lietz (DePaul U.), Gordon Lind (Utah State U.), S. Livingston
(U. of Wisconsin, Milwaukee), Jorge Lopez (U. of Texas, El Paso), Elihu Lubkin (U. of
Wisconsin, Milwaukee), Robert Luke (Boise State U.), David Lynch (Iowa State U.),
Michael Lysak (San Bernardino Valley C.), Jeffrey Mallow (Loyola U.), Robert Mania
(Kentucky State U.), Robert Marchina (U. of Memphis), David Markowitz (U. of
Connecticut), Philip Matheson (Utah Valley U.), R. J. Maurer, Oren Maxwell (Florida
International U.), Joseph L. McCauley (U. of Houston), T. K. McCubbin, Jr. (Pennsylvania
State U.), Charles McFarland (U. of Missouri at Rolla), James Mcguire (Tulane U.),
Lawrence McIntyre (U. of Arizona), Fredric Messing (Carnegie Mellon U.), Thomas
Meyer (Texas A&M U.), Andre Mirabelli (St. Peter’s C., New Jersey), Herbert Muether
PreFace
(SUNY, Stony Brook), Jack Munsee (California State U., Long Beach), Lorenzo Narducci
(Drexel U.), Van E. Neie (Purdue U.), Forrest Newman (Sacramento City C.), David A.
Nordling (U.S. Naval Academy), Benedict Oh (Pennsylvania State U.), L. O. Olsen,
Michael Ottinger (Missouri Western State U.), Russell Palma (Minnesota State U.,
Mankato), Jim Pannell (DeVry Institute of Technology), Neeti Parashar (Purdue U.,
Calumet), W. F. Parks (U. of Missouri), Robert Paulson (California State U., Chico), Jerry
Peacher (U. of Missouri at Rolla), Arnold Perlmutter (U. of Miami), Lennart Peterson
(U. of Florida), R. J. Peterson (U. of Colorado, Boulder), R. Pinkston, Ronald Poling
(U. of Minnesota), Yuri Popov (U. of Michigan), J. G. Potter, C. W. Price (Millersville U.),
Francis Prosser (U. of Kansas), Shelden H. Radin, Roberto Ramos (Drexel U.),
Michael Rapport (Anne Arundel Comm. C.), R. Resnick, James A. Richards, Jr., John S.
Risley (North Carolina State U.), Francesc Roig (UC, Santa Barbara), T. L. Rokoske,
Richard Roth (Eastern Michigan U.), Carl Rotter (U. of West Virginia), S. Clark Rowland
(Andrews U.), Rajarshi Roy (Georgia Institute of Technology), Russell A. Roy (Santa Fe
Comm. C.), Desi Saludes (Hillsborough Comm. C.), Thomas Sandin (North Carolina A&T
State U.), Dhiraj Sardar (U. of Texas, San Antonio), Tumer Sayman (Eastern Michigan U.),
Bruce Schumm (UC, Santa Cruz), Melvin Schwartz (St. John’s U.), F. A. Scott, L. W.
Seagondollar, Paul Shand (U. of Northern Iowa), Stan Shepherd (Pennsylvania State U.),
Douglas Sherman (San Jose State U.), Bruce Sherwood (Carnegie Mellon U.), Hugh Siefkin
(Greenville C.), Christopher Sirola (U. of Southern Mississippi), Tomasz Skwarnicki
(Syracuse U.), C. P. Slichter, Jason Slinker (U. of Texas, Dallas), Charles W. Smith (U. of
Maine, Orono), Malcolm Smith (U. of Lowell), Ross Spencer (Brigham Young U.), Julien
Sprott (U. of Wisconsin), Victor Stanionis (Iona C.), James Stith (American Institute of
Physics), Chuck Stone (North Carolina A&T State U.), Edward Strother (Florida Institute of
Technology), Conley Stutz (Bradley U.), Albert Stwertka (U.S. Merchant Marine Academy),
Kenneth Szpara-DeNisco (Harrisburg Area Comm. C.), Devki Talwar (Indiana U. of
Pennsylvania), Fiorella Terenzi (Florida International U.), Martin Tiersten (CUNY, City C.),
David Toot (Alfred U.), Greg Trayling (Rochester Institute of Technology), Somdev
Tyagi (Drexel U.), Matthew Vannette (Saginaw Valley State U.), Eswara Venugopal
(U. of Detroit, Mercy), F. Verbrugge, Helmut Vogel (Carnegie Mellon U.), Aaron Warren
(Purdue U., North Central), Robert Webb (Texas A&M U.), Thomas Weber (Iowa State U.),
M. Russell Wehr (Pennsylvania State U.), Robert Weidman (Michigan Technical U.),
Dan Whalen (UC, San Diego), Lester V. Whitney, Thomas Wiggins (Pennsylvania State U.),
Robyn Wilde (Oregon Institute of Technology), David Willey (U. of Pittsburgh,
Johnstown), George Williams (U. of Utah), John Williams (Auburn U.), Stanley Williams
(Iowa State U.), Jack Willis, Suzanne Willis (Northern Illinois U.), Robert Wilson (San
Bernardino Valley C.), L. Wolfenstein, James Wood (Palm Beach Junior C.), Lowell
Wood (U. of Houston), R. E. Worley, D. H. Ziebell (Manatee Comm. C.), George O.
Zimmerman (Boston U.)
In addition, I would like to thank my past and present colleagues at UCSB,
including Rob Geller, Carl Gwinn, Al Nash, Elisabeth Nicol, and Francesc Roig,
for their wholehearted support and for many helpful discussions. I owe a special
debt of gratitude to my early teachers Willa Ramsay, Peter Zimmerman, William
Little, Alan Schwettman, and Dirk Walecka for showing me what clear and engaging physics teaching is all about, and to Stuart Johnson for inviting me to become
a coauthor of University Physics beginning with the Ninth Edition. Special
acknowledgments go out to Lewis Ford for creating a wealth of new problems for
this edition, including the new category of DATA problems; to Wayne Anderson,
who carefully reviewed all of the problems and solved them, along with Forrest
Newman and Michael Ottinger; and to Elizabeth George, who provided most of
the new category of Passage Problems. A particular tip of the hat goes to Tom
Sandin for his numerous contributions to the end-of-chapter problems, including
carefully checking all problems and writing new ones. Hats off as well and a tremendous reception to Linda Hirst for contributing a number of ideas that became
new Application features in this edition. I want to express special thanks to the
editorial staff at Pearson: to Nancy Whilton for her editorial vision; to Karen Karlin
for her keen eye and careful development of this edition; to Charles Hibbard
for his careful reading of the page proofs; and to Beth Collins, Katie Conley,
Sarah Kaubisch, Eric Schrader, and Cindy Johnson for keeping the editorial and
xvii
xviii
PreFace
production pipelines flowing. Most of all, I want to express my gratitude and
love to my wife, Caroline, to whom I dedicate my contribution to this book. Hey,
Caroline, the new edition’s done at last—let’s go flying!
pleAse tell Me WhAt yOu think!
I welcome communications from students and professors, especially concerning
errors or deficiencies that you find in this edition. The late Hugh Young and I have
devoted a lot of time and effort to writing the best book we know how to write, and
I hope it will help as you teach and learn physics. In turn, you can help me by letting me know what still needs to be improved! Please feel free to contact me either
electronically or by ordinary mail. Your comments will be greatly appreciated.
August 2014
Roger A. Freedman
Department of Physics
University of California, Santa Barbara
Santa Barbara, CA 93106-9530
airboy@physics.ucsb.edu
http://www.physics.ucsb.edu/~airboy/
Twitter: @RogerFreedman
DetAiLeD ContentS
MechAnics
1
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
1.9
1.10
2
2.1
2.2
2.3
2.4
2.5
2.6
3
3.1
3.2
3.3
3.4
3.5
4
4.1
4.2
4.3
4.4
4.5
4.6
UNITS, PHYSICAL QUANTITIES,
AND VECTORS
The Nature of Physics
Solving Physics Problems
Standards and Units
Using and Converting Units
Uncertainty and Significant Figures
Estimates and Orders of Magnitude
Vectors and Vector Addition
Components of Vectors
Unit Vectors
Products of Vectors
Summary
Questions/Exercises/Problems
1
2
2
4
6
8
10
10
14
18
19
25
27
5
5.1
5.2
MOTION ALONG
A STRAIGHT LINE
34
Displacement, Time, and Average
Velocity
Instantaneous Velocity
Average and Instantaneous Acceleration
Motion with Constant Acceleration
Freely Falling Bodies
Velocity and Position by Integration
Summary
Questions/Exercises/Problems
34
37
40
45
50
53
56
57
MOTION IN TWO
OR THREE DIMENSIONS
67
Position and Velocity Vectors
The Acceleration Vector
Projectile Motion
Motion in a Circle
Relative Velocity
Summary
Questions/Exercises/Problems
67
70
75
82
86
91
92
NEWTON’S LAWS OF MOTION
101
Force and Interactions
Newton’s First Law
Newton’s Second Law
Mass and Weight
Newton’s Third Law
Free-Body Diagrams
Summary
Questions/Exercises/Problems
102
105
108
114
116
120
121
123
5.3
5.4
5.5
6
6.1
6.2
6.3
6.4
7
7.1
7.2
7.3
7.4
7.5
8
8.1
8.2
APPLYING NEWTON’S LAWS
Using Newton’s First Law:
Particles in Equilibrium
Using Newton’s Second Law:
Dynamics of Particles
Friction Forces
Dynamics of Circular Motion
The Fundamental Forces of Nature
Summary
Questions/Exercises/Problems
130
130
135
142
150
155
157
159
WORK AND KINETIC ENERGY
172
Work
Kinetic Energy and the Work–Energy
Theorem
Work and Energy with Varying Forces
Power
Summary
Questions/Exercises/Problems
173
177
183
189
192
193
POTENTIAL ENERGY
AND ENERGY CONSERVATION
203
Gravitational Potential Energy
Elastic Potential Energy
Conservative and Nonconservative Forces
Force and Potential Energy
Energy Diagrams
Summary
Questions/Exercises/Problems
203
212
217
221
224
226
227
MOMENTUM, IMPULSE,
AND COLLISIONS
237
Momentum and Impulse
Conservation of Momentum
238
243
xx
detailed contentS
8.3
8.4
8.5
8.6
9
Momentum Conservation and Collisions
Elastic Collisions
Center of Mass
Rocket Propulsion
Summary
Questions/Exercises/Problems
247
251
254
258
261
262
12
FLUID MECHANICS
369
12.1
12.2
12.3
12.4
12.5
12.6
Gases, Liquids, and Density
Pressure in a Fluid
Buoyancy
Fluid Flow
Bernoulli’s Equation
Viscosity and Turbulence
Summary
Questions/Exercises/Problems
369
371
376
379
381
385
388
389
13
GRAVITATION
398
13.1
13.2
13.3
13.4
13.5
13.6
13.7
13.8
Newton’s Law of Gravitation
Weight
Gravitational Potential Energy
The Motion of Satellites
Kepler’s Laws and the Motion of Planets
Spherical Mass Distributions
Apparent Weight and the Earth’s Rotation
Black Holes
Summary
Questions/Exercises/Problems
398
402
405
407
410
414
417
419
423
424
ROTATION OF RIGID BODIES
273
Angular Velocity and Acceleration
Rotation with Constant Angular
Acceleration
Relating Linear and Angular Kinematics
Energy in Rotational Motion
Parallel-Axis Theorem
Moment-of-Inertia Calculations
Summary
Questions/Exercises/Problems
273
14
PERIODIC MOTION
433
278
280
283
288
289
292
293
14.1
14.2
14.3
14.4
14.5
14.6
14.7
14.8
10
DYNAMICS
OF ROTATIONAL MOTION
303
Describing Oscillation
Simple Harmonic Motion
Energy in Simple Harmonic Motion
Applications of Simple Harmonic Motion
The Simple Pendulum
The Physical Pendulum
Damped Oscillations
Forced Oscillations and Resonance
Summary
Questions/Exercises/Problems
433
435
442
446
450
451
453
455
457
459
10.1
10.2
Torque
Torque and Angular Acceleration for a
Rigid Body
Rigid-Body Rotation About a Moving Axis
Work and Power in Rotational Motion
Angular Momentum
Conservation of Angular Momentum
Gyroscopes and Precession
Summary
Questions/Exercises/Problems
9.1
9.2
9.3
9.4
9.5
9.6
10.3
10.4
10.5
10.6
10.7
303
306
309
315
317
320
322
326
327
11
EQUILIBRIUM AND ELASTICITY 339
11.1
11.2
11.3
11.4
11.5
Conditions for Equilibrium
Center of Gravity
Solving Rigid-Body Equilibrium Problems
Stress, Strain, and Elastic Moduli
Elasticity and Plasticity
Summary
Questions/Exercises/Problems
340
340
343
347
353
354
356
WAves/AcOustics
15
MECHANICAL WAVES
468
15.1
15.2
15.3
15.4
15.5
15.6
Types of Mechanical Waves
Periodic Waves
Mathematical Description of a Wave
Speed of a Transverse Wave
Energy in Wave Motion
Wave Interference, Boundary Conditions,
and Superposition
Standing Waves on a String
Normal Modes of a String
Summary
Questions/Exercises/Problems
468
470
473
478
482
16
SOUND AND HEARING
505
16.1
16.2
Sound Waves
Speed of Sound Waves
505
510
15.7
15.8
485
487
491
495
496
detailed contentS
16.3
16.4
16.5
16.6
16.7
16.8
16.9
Sound Intensity
Standing Sound Waves and Normal Modes
Resonance and Sound
Interference of Waves
Beats
The Doppler Effect
Shock Waves
Summary
Questions/Exercises/Problems
xxi
514
518
522
524
526
528
533
535
537
therMOdynAMics
17
TEMPERATURE AND HEAT
545
17.1
17.2
17.3
17.4
17.5
17.6
17.7
Temperature and Thermal Equilibrium
Thermometers and Temperature Scales
Gas Thermometers and the Kelvin Scale
Thermal Expansion
Quantity of Heat
Calorimetry and Phase Changes
Mechanisms of Heat Transfer
Summary
Questions/Exercises/Problems
545
547
548
551
556
559
565
572
573
18
THERMAL PROPERTIES
OF MATTER
584
21
18.1
18.2
18.3
18.4
18.5
18.6
Equations of State
Molecular Properties of Matter
Kinetic-Molecular Model of an Ideal Gas
Heat Capacities
Molecular Speeds
Phases of Matter
Summary
Questions/Exercises/Problems
585
590
593
599
602
604
607
609
ELECTRIC CHARGE
AND ELECTRIC FIELD
683
21.1
21.2
21.3
21.4
21.5
21.6
21.7
19
THE FIRST LAW
OF THERMODYNAMICS
Electric Charge
Conductors, Insulators, and Induced Charges
Coulomb’s Law
Electric Field and Electric Forces
Electric-Field Calculations
Electric Field Lines
Electric Dipoles
Summary
Questions/Exercises/Problems
684
687
690
695
699
705
706
711
712
618
19.1
19.2
19.3
19.4
Thermodynamic Systems
Work Done During Volume Changes
Paths Between Thermodynamic States
Internal Energy and the First Law of
Thermodynamics
Kinds of Thermodynamic Processes
Internal Energy of an Ideal Gas
Heat Capacities of an Ideal Gas
Adiabatic Processes for an Ideal Gas
Summary
Questions/Exercises/Problems
22
GAUSS’S LAW
722
22.1
22.2
22.3
22.4
22.5
Charge and Electric Flux
Calculating Electric Flux
Gauss’s Law
Applications of Gauss’s Law
Charges on Conductors
Summary
Questions/Exercises/Problems
722
725
729
733
738
743
744
23
ELECTRIC POTENTIAL
752
23.1
23.2
23.3
23.4
23.5
Electric Potential Energy
Electric Potential
Calculating Electric Potential
Equipotential Surfaces
Potential Gradient
Summary
Questions/Exercises/Problems
752
759
765
769
771
775
776
19.5
19.6
19.7
19.8
618
620
622
623
628
630
631
634
637
638
20
THE SECOND LAW
OF THERMODYNAMICS
647
20.1
20.2
20.3
Directions of Thermodynamic Processes
Heat Engines
Internal-Combustion Engines
647
649
652
20.4
20.5
20.6
20.7
20.8
Refrigerators
The Second Law of Thermodynamics
The Carnot Cycle
Entropy
Microscopic Interpretation of Entropy
Summary
Questions/Exercises/Problems
654
656
658
664
670
674
676
electrOMAgnetisM
xxii
24
24.1
24.2
24.3
24.4
24.5
24.6
detailed contentS
CAPACITANCE
AND DIELECTRICS
Capacitors and Capacitance
Capacitors in Series and Parallel
Energy Storage in Capacitors and
Electric-Field Energy
Dielectrics
Molecular Model of Induced Charge
Gauss’s Law in Dielectrics
Summary
Questions/Exercises/Problems
27.5
785
786
790
794
797
803
805
806
808
25
CURRENT, RESISTANCE,
AND ELECTROMOTIVE FORCE
816
25.1
25.2
25.3
25.4
25.5
25.6
Current
Resistivity
Resistance
Electromotive Force and Circuits
Energy and Power in Electric Circuits
Theory of Metallic Conduction
Summary
Questions/Exercises/Problems
817
820
823
826
832
836
839
840
27.6
27.7
27.8
27.9
DIRECT-CURRENT CIRCUITS
848
26.1
26.2
26.3
26.4
26.5
Resistors in Series and Parallel
Kirchhoff’s Rules
Electrical Measuring Instruments
R-C Circuits
Power Distribution Systems
Summary
Questions/Exercises/Problems
848
853
858
862
867
871
872
27
MAGNETIC FIELD AND
MAGNETIC FORCES
881
27.1
27.2
27.3
27.4
Magnetism
Magnetic Field
Magnetic Field Lines and Magnetic Flux
Motion of Charged Particles
in a Magnetic Field
881
883
887
890
894
896
900
905
907
909
911
28
SOURCES OF MAGNETIC FIELD
921
28.1
28.2
28.3
Magnetic Field of a Moving Charge
Magnetic Field of a Current Element
Magnetic Field of a Straight
Current-Carrying Conductor
Force Between Parallel Conductors
Magnetic Field of a Circular
Current Loop
Ampere’s Law
Applications of Ampere’s Law
Magnetic Materials
Summary
Questions/Exercises/Problems
921
924
28.4
28.5
28.6
28.7
28.8
926
929
930
933
936
939
945
947
29
ELECTROMAGNETIC
INDUCTION
29.1
29.2
29.3
29.4
29.5
29.6
29.7
Induction Experiments
Faraday’s Law
Lenz’s Law
Motional Electromotive Force
Induced Electric Fields
Eddy Currents
Displacement Current and Maxwell’s
Equations
Superconductivity
Summary
Questions/Exercises/Problems
956
957
965
967
969
972
30
INDUCTANCE
990
30.1
30.2
30.3
30.4
30.5
30.6
Mutual Inductance
Self-Inductance and Inductors
Magnetic-Field Energy
The R-L Circuit
The L-C Circuit
The L-R-C Series Circuit
Summary
Questions/Exercises/Problems
990
994
997
1000
1004
1008
1011
1012
31
ALTERNATING CURRENT
1020
31.1
31.2
31.3
31.4
Phasors and Alternating Currents
Resistance and Reactance
The L-R-C Series Circuit
Power in Alternating-Current Circuits
1020
1023
1028
1033
29.8
26
Applications of Motion of
Charged Particles
Magnetic Force on a Current-Carrying
Conductor
Force and Torque on a Current Loop
The Direct-Current Motor
The Hall Effect
Summary
Questions/Exercises/Problems
955
973
977
979
980
detailed contentS
31.5
31.6
Resonance in Alternating-Current
Circuits
Transformers
Summary
Questions/Exercises/Problems
1036
1038
1042
1043
32
ELECTROMAGNETIC WAVES
1050
32.1
Maxwell’s Equations and
Electromagnetic Waves
Plane Electromagnetic Waves
and the Speed of Light
Sinusoidal Electromagnetic Waves
Energy and Momentum
in Electromagnetic Waves
Standing Electromagnetic Waves
Summary
Questions/Exercises/Problems
32.2
32.3
32.4
32.5
1051
1054
1059
1063
1068
1071
1072
Optics
33
THE NATURE AND
PROPAGATION OF LIGHT
1078
33.1
33.2
33.3
33.4
33.5
33.6
33.7
The Nature of Light
Reflection and Refraction
Total Internal Reflection
Dispersion
Polarization
Scattering of Light
Huygens’s Principle
Summary
Questions/Exercises/Problems
1078
1080
1086
1089
1091
1099
1100
1102
1104
34
GEOMETRIC OPTICS
1111
34.1
Reflection and Refraction at a
Plane Surface
Reflection at a Spherical Surface
Refraction at a Spherical Surface
Thin Lenses
Cameras
The Eye
The Magnifier
Microscopes and Telescopes
Summary
Questions/Exercises/Problems
1111
1115
1123
1128
1136
1139
1143
1144
1149
1151
35
INTERFERENCE
1160
35.1
35.2
35.3
35.4
35.5
Interference and Coherent Sources
Two-Source Interference of Light
Intensity in Interference Patterns
Interference in Thin Films
The Michelson Interferometer
Summary
Questions/Exercises/Problems
1160
1164
1167
1171
1176
1178
1179
34.2
34.3
34.4
34.5
34.6
34.7
34.8
xxiii
36
DIFFRACTION
1186
36.1
36.2
36.3
36.4
36.5
36.6
36.7
36.8
Fresnel and Fraunhofer Diffraction
Diffraction from a Single Slit
Intensity in the Single-Slit Pattern
Multiple Slits
The Diffraction Grating
X-Ray Diffraction
Circular Apertures and Resolving Power
Holography
Summary
Questions/Exercises/Problems
1186
1188
1191
1195
1197
1201
1204
1207
1209
1210
MOdern physics
37
RELATIVITY
1218
37.1
37.2
37.3
37.4
37.5
37.6
Invariance of Physical Laws
Relativity of Simultaneity
Relativity of Time Intervals
Relativity of Length
The Lorentz Transformations
The Doppler Effect for
Electromagnetic Waves
Relativistic Momentum
Relativistic Work and Energy
Newtonian Mechanics and Relativity
Summary
Questions/Exercises/Problems
1218
1221
1223
1228
1232
1236
1238
1240
1244
1245
1247
38
PHOTONS: LIGHT WAVES
BEHAVING AS PARTICLES
1254
38.1
Light Absorbed as Photons:
The Photoelectric Effect
Light Emitted as Photons:
X-Ray Production
Light Scattered as Photons: Compton
Scattering and Pair Production
Wave–Particle Duality, Probability,
and Uncertainty
Summary
Questions/Exercises/Problems
37.7
37.8
37.9
38.2
38.3
38.4
1254
1260
1263
1266
1273
1274
xxiv
detailed contentS
T-bacteriophage viruses
39.1
39.2
39.3
39.4
39.5
39.6
PARTICLES BEHAVING
AS WAVES
Electron Waves
The Nuclear Atom and Atomic Spectra
Energy Levels and the Bohr Model
of the Atom
The Laser
Continuous Spectra
The Uncertainty Principle Revisited
Summary
Questions/Exercises/Problems
1279
1279
1285
1290
1300
1303
1308
1311
1313
40
QUANTUM MECHANICS I:
WAVE FUNCTIONS
1321
40.1
Wave Functions and the One-Dimensional
Schrödinger Equation
Particle in a Box
Potential Wells
Potential Barriers and Tunneling
The Harmonic Oscillator
Measurement in Quantum Mechanics
Summary
Questions/Exercises/Problems
1321
1331
1336
1340
1343
1348
1351
1353
41
QUANTUM MECHANICS II:
ATOMIC STRUCTURE
1360
41.1
The Schrödinger Equation in
Three Dimensions
Particle in a Three-Dimensional Box
The Hydrogen Atom
The Zeeman Effect
Electron Spin
Many-Electron Atoms and the Exclusion
Principle
X-Ray Spectra
Quantum Entanglement
Summary
Questions/Exercises/Problems
40.2
40.3
40.4
40.5
40.6
41.2
41.3
41.4
41.5
41.6
41.7
41.8
MOLECULES
AND CONDENSED MATTER
1407
42.1
42.2
42.3
42.4
42.5
42.6
42.7
42.8
Types of Molecular Bonds
Molecular Spectra
Structure of Solids
Energy Bands
Free-Electron Model of Metals
Semiconductors
Semiconductor Devices
Superconductivity
Summary
Questions/Exercises/Problems
1407
1410
1414
1418
1420
1424
1427
1432
1432
1434
43
NUCLEAR PHYSICS
1440
43.1
43.2
Properties of Nuclei
Nuclear Binding and Nuclear
Structure
Nuclear Stability and Radioactivity
Activities and Half-Lives
Biological Effects of Radiation
Nuclear Reactions
Nuclear Fission
Nuclear Fusion
Summary
Questions/Exercises/Problems
1440
1446
1450
1457
1461
1464
1466
1470
1473
1474
44
PARTICLE PHYSICS
AND COSMOLOGY
1481
44.1
44.2
44.3
44.4
44.5
44.6
44.7
Fundamental Particles—A History
Particle Accelerators and Detectors
Particles and Interactions
Quarks and Gluons
The Standard Model and Beyond
The Expanding Universe
The Beginning of Time
Summary
Questions/Exercises/Problems
1481
1486
1490
1496
1500
1502
1509
1517
1519
100 nm = 0.1 Mm
Viral
DNA
39
42
1360
1362
1367
1375
1378
1385
1392
1395
1399
1401
43.3
43.4
43.5
43.6
43.7
43.8
Appendices
A
B
C
D
E
F
The International System of Units
Useful Mathematical Relations
The Greek Alphabet
Periodic Table of the Elements
Unit Conversion Factors
Numerical Constants
A-1
A-3
A-4
A-5
A-6
A-7
Answers to Odd-Numbered Problems
Credits
Index
A-9
C-1
I-1
?
Tornadoes are spawned by
severe thunderstorms, so
being able to predict the path
of thunderstorms is essential.
If a thunderstorm is moving at
15 km/h in a direction 37° north
of east, how far north does the
thunderstorm move in 2.0 h?
(i) 30 km; (ii) 24 km; (iii) 18 km;
(iv) 12 km; (v) 9 km.
1 Units, Physical
QUantities,
and Vectors
Learning goaLs
Looking forward at …
1.1 What a physical theory is.
1.2 The four steps you can use to solve any
physics problem.
1.3 Three fundamental quantities of physics
1.4
1.5
1.6
1.7
1.8
1.9
1.10
and the units physicists use to measure
them.
How to work with units in your
calculations.
How to keep track of significant figures in
your calculations.
How to make rough, order-of-magnitude
estimates.
The difference between scalars and
vectors, and how to add and subtract
vectors graphically.
What the components of a vector are and
how to use them in calculations.
What unit vectors are and how to use them
with components to describe vectors.
Two ways to multiply vectors: the scalar
(dot) product and the vector (cross)
product.
P
hysics is one of the most fundamental of the sciences. Scientists of all
disciplines use the ideas of physics, including chemists who study the
structure of molecules, paleontologists who try to reconstruct how
dinosaurs walked, and climatologists who study how human activities affect the
atmosphere and oceans. Physics is also the foundation of all engineering and
technology. No engineer could design a flat-screen TV, a prosthetic leg, or even a
better mousetrap without first understanding the basic laws of physics.
The study of physics is also an adventure. You will find it challenging, sometimes frustrating, occasionally painful, and often richly rewarding. If you’ve ever
wondered why the sky is blue, how radio waves can travel through empty space,
or how a satellite stays in orbit, you can find the answers by using fundamental
physics. You will come to see physics as a towering achievement of the human
intellect in its quest to understand our world and ourselves.
In this opening chapter, we’ll go over some important preliminaries that we’ll
need throughout our study. We’ll discuss the nature of physical theory and the
use of idealized models to represent physical systems. We’ll introduce the systems of units used to describe physical quantities and discuss ways to describe
the accuracy of a number. We’ll look at examples of problems for which we can’t
(or don’t want to) find a precise answer, but for which rough estimates can be
useful and interesting. Finally, we’ll study several aspects of vectors and vector
algebra. We’ll need vectors throughout our study of physics to help us describe
and analyze physical quantities, such as velocity and force, that have direction as
well as magnitude.
1
2
Chapter 1 Units, physical Quantities, and Vectors
1.1 The NaTure of Physics
Physics is an experimental science. Physicists observe the phenomena of nature
and try to find patterns that relate these phenomena. These patterns are called
physical theories or, when they are very well established and widely used, physical laws or principles.
cauTioN The meaning of “theory” A theory is not just a random thought or an unproven
concept. Rather, a theory is an explanation of natural phenomena based on observation
and accepted fundamental principles. An example is the well-established theory of biological evolution, which is the result of extensive research and observation by generations
of biologists. ❙
1.1 Two research laboratories.
(a) According to legend, Galileo investigated
falling objects by dropping them from the
Leaning Tower of Pisa, Italy, ...
... and he studied pendulum motion
by observing the swinging chandelier
in the adjacent cathedral.
(b) The Planck spacecraft is designed to study
the faint electromagnetic radiation left over
from the Big Bang 13.8 billion years ago.
To develop a physical theory, a physicist has to learn to ask appropriate questions, design experiments to try to answer the questions, and draw appropriate
conclusions from the results. Figure 1.1 shows two important facilities used for
physics experiments.
Legend has it that Galileo Galilei (1564–1642) dropped light and heavy objects from the top of the Leaning Tower of Pisa (Fig. 1.1a) to find out whether
their rates of fall were different. From examining the results of his experiments
(which were actually much more sophisticated than in the legend), he made the
inductive leap to the principle, or theory, that the acceleration of a falling object
is independent of its weight.
The development of physical theories such as Galileo’s often takes an indirect
path, with blind alleys, wrong guesses, and the discarding of unsuccessful theories in favor of more promising ones. Physics is not simply a collection of facts
and principles; it is also the process by which we arrive at general principles that
describe how the physical universe behaves.
No theory is ever regarded as the final or ultimate truth. The possibility always exists that new observations will require that a theory be revised or discarded. It is in the nature of physical theory that we can disprove a theory by
finding behavior that is inconsistent with it, but we can never prove that a theory
is always correct.
Getting back to Galileo, suppose we drop a feather and a cannonball. They
certainly do not fall at the same rate. This does not mean that Galileo was wrong;
it means that his theory was incomplete. If we drop the feather and the cannonball in a vacuum to eliminate the effects of the air, then they do fall at the same
rate. Galileo’s theory has a range of validity: It applies only to objects for which
the force exerted by the air (due to air resistance and buoyancy) is much less than
the weight. Objects like feathers or parachutes are clearly outside this range.
1.2 solviNg Physics Problems
These technicians are reflected in
the spacecraft’s light-gathering
mirror during pre-launch testing.
University Physics 14e
Young/Freedman
Benjamin Cummings
At some point in their studies, almost all physics students find themselves
thinking, “I understand the concepts, but I just can’t solve the problems.” But in
physics, truly understanding a concept means being able to apply it to a variety of
problems. Learning how to solve problems is absolutely essential; you don’t know
physics unless you can do physics.
How do you learn to solve physics problems? In every chapter of this book
you will find Problem-Solving Strategies that offer techniques for setting up and
solving problems efficiently and accurately. Following each Problem-Solving
Strategy are one or more worked Examples that show these techniques in action.
(The Problem-Solving Strategies will also steer you away from some incorrect
techniques that you may be tempted to use.) You’ll also find additional examples
that aren’t associated with a particular Problem-Solving Strategy. In addition, at
the end of each chapter you’ll find a Bridging Problem that uses more than one of
1.2 Solving Physics Problems
3
the key ideas from the chapter. Study these strategies and problems carefully, and
work through each example for yourself on a piece of paper.
Different techniques are useful for solving different kinds of physics problems, which is why this book offers dozens of Problem-Solving Strategies. No
matter what kind of problem you’re dealing with, however, there are certain key
steps that you’ll always follow. (These same steps are equally useful for problems
in math, engineering, chemistry, and many other fields.) In this book we’ve organized these steps into four stages of solving a problem.
All of the Problem-Solving Strategies and Examples in this book will follow
these four steps. (In some cases we will combine the first two or three steps.) We
encourage you to follow these same steps when you solve problems yourself. You
may find it useful to remember the acronym I SEE—short for Identify, Set up,
Execute, and Evaluate.
Problem-Solving STraTegy 1.1
Solving PhySicS ProblemS
idenTify the relevant concepts: Use the physical conditions stated
in the problem to help you decide which physics concepts are relevant. Identify the target variables of the problem—that is, the
quantities whose values you’re trying to find, such as the speed at
which a projectile hits the ground, the intensity of a sound made
by a siren, or the size of an image made by a lens. Identify the
known quantities, as stated or implied in the problem. This step is
essential whether the problem asks for an algebraic expression or
a numerical answer.
seT uP the problem: Given the concepts you have identified, the
known quantities, and the target variables, choose the equations
that you’ll use to solve the problem and decide how you’ll use
them. Make sure that the variables you have identified correlate
exactly with those in the equations. If appropriate, draw a sketch
of the situation described in the problem. (Graph paper, ruler, protractor, and compass will help you make clear, useful sketches.)
As best you can, estimate what your results will be and, as appropriate, predict what the physical behavior of a system will be.
The worked examples in this book include tips on how to make
these kinds of estimates and predictions. If this seems challenging, don’t worry—you’ll get better with practice!
execuTe the solution: This is where you “do the math.” Study the
worked examples to see what’s involved in this step.
evaLuaTe your answer: Compare your answer with your estimates, and reconsider things if there’s a discrepancy. If your answer includes an algebraic expression, assure yourself that it
correctly represents what would happen if the variables in it had
very large or very small values. For future reference, make note of
any answer that represents a quantity of particular significance.
Ask yourself how you might answer a more general or more difficult version of the problem you have just solved.
idealized models
In everyday conversation we use the word “model” to mean either a small-scale
replica, such as a model railroad, or a person who displays articles of clothing
(or the absence thereof). In physics a model is a simplified version of a physical
system that would be too complicated to analyze in full detail.
For example, suppose we want to analyze the motion of a thrown baseball
(Fig. 1.2a). How complicated is this problem? The ball is not a perfect sphere
(it has raised seams), and it spins as it moves through the air. Air resistance and
wind influence its motion, the ball’s weight varies a little as its altitude changes,
and so on. If we try to include all these things, the analysis gets hopelessly complicated. Instead, we invent a simplified version of the problem. We ignore the
size and shape of the ball by representing it as a point object, or particle. We
ignore air resistance by making the ball move in a vacuum, and we make the
weight constant. Now we have a problem that is simple enough to deal with
(Fig. 1.2b). We will analyze this model in detail in Chapter 3.
We have to overlook quite a few minor effects to make an idealized model,
but we must be careful not to neglect too much. If we ignore the effects of gravity completely, then our model predicts that when we throw the ball up, it will go
in a straight line and disappear into space. A useful model simplifies a problem
enough to make it manageable, yet keeps its essential features.
1.2 To simplify the analysis of (a) a baseball in flight, we use (b) an idealized
model.
(a) A real baseball in flight
Baseball spins and has a complex shape.
Air resistance and
wind exert forces
on the ball.
Direction of
motion
Gravitational force on ball
depends on altitude.
(b) An idealized model of the baseball
Treat the baseball as a point object (particle).
No air resistance.
Gravitational force
on ball is constant.
Direction of
motion
4
ChaPter 1 Units, Physical Quantities, and Vectors
The validity of the predictions we make using a model is limited by the validity of the model. For example, Galileo’s prediction about falling objects (see
Section 1.1) corresponds to an idealized model that does not include the effects
of air resistance. This model works fairly well for a dropped cannonball, but not
so well for a feather.
Idealized models play a crucial role throughout this book. Watch for them in
discussions of physical theories and their applications to specific problems.
1.3 sTandards and uniTs
1.3 The measurements used to determine
(a) the duration of a second and (b) the
length of a meter. These measurements
are useful for setting standards because
they give the same results no matter where
they are made.
(a) Measuring the second
Microwave radiation with a frequency of
exactly 9,192,631,770 cycles per second ...
Outermost
electron
Cesium-133
atom
... causes the outermost electron of a
cesium-133 atom to reverse its spin direction.
As we learned in Section 1.1, physics is an experimental science. Experiments
require measurements, and we generally use numbers to describe the results
of measurements. Any number that is used to describe a physical phenomenon
quantitatively is called a physical quantity. For example, two physical quantities that describe you are your weight and your height. Some physical quantities
are so fundamental that we can define them only by describing how to measure
them. Such a definition is called an operational definition. Two examples are
measuring a distance by using a ruler and measuring a time interval by using
a stopwatch. In other cases we define a physical quantity by describing how to
calculate it from other quantities that we can measure. Thus we might define the
average speed of a moving object as the distance traveled (measured with a ruler)
divided by the time of travel (measured with a stopwatch).
When we measure a quantity, we always compare it with some reference standard. When we say that a Ferrari 458 Italia is 4.53 meters long, we mean that it
is 4.53 times as long as a meter stick, which we define to be 1 meter long. Such
a standard defines a unit of the quantity. The meter is a unit of distance, and the
second is a unit of time. When we use a number to describe a physical quantity,
we must always specify the unit that we are using; to describe a distance as simply “4.53” wouldn’t mean anything.
To make accurate, reliable measurements, we need units of measurement
that do not change and that can be duplicated by observers in various locations.
The system of units used by scientists and engineers around the world is commonly called “the metric system,” but since 1960 it has been known officially as
the International System, or SI (the abbreviation for its French name, Système
International). Appendix A gives a list of all SI units as well as definitions of the
most fundamental units.
Time
Cesium-133
atom
An atomic clock uses this phenomenon to tune
microwaves to this exact frequency. It then
counts 1 second for each 9,192,631,770 cycles.
Length
(b) Measuring the meter
0:00 s
Light
source
0:01 s
Light travels exactly
299,792,458 m in 1 s.
From 1889 until 1967, the unit of time was defined as a certain fraction of the
mean solar day, the average time between successive arrivals of the sun at its
highest point in the sky. The present standard, adopted in 1967, is much more
precise. It is based on an atomic clock, which uses the energy difference between
the two lowest energy states of the cesium atom (133Cs). When bombarded by
microwaves of precisely the proper frequency, cesium atoms undergo a transition
from one of these states to the other. One second (abbreviated s) is defined as the
time required for 9,192,631,770 cycles of this microwave radiation (Fig. 1.3a).
In 1960 an atomic standard for the meter was also established, using the
wavelength of the orange-red light emitted by excited atoms of krypton
186Kr2. From this length standard, the speed of light in vacuum was measured to
be 299,792,458 m>s. In November 1983, the length standard was changed again
so that the speed of light in vacuum was defined to be precisely 299,792,458 m>s.
1.3 Standards and Units
Hence the new definition of the meter (abbreviated m) is the distance that light
travels in vacuum in 1>299,792,458 second (Fig. 1.3b). This modern definition
provides a much more precise standard of length than the one based on a wavelength of light.
1.4 The international standard kilogram
is the metal object carefully enclosed
within these nested glass containers.
Mass
The standard of mass, the kilogram (abbreviated kg), is defined to be the mass of
a particular cylinder of platinum–iridium alloy kept at the International Bureau
of Weights and Measures at Sèvres, near Paris (Fig. 1.4). An atomic standard
of mass would be more fundamental, but at present we cannot measure masses
on an atomic scale with as much accuracy as on a macroscopic scale. The gram
(which is not a fundamental unit) is 0.001 kilogram.
Other derived units can be formed from the fundamental units. For example,
the units of speed are meters per second, or m>s; these are the units of length (m)
divided by the units of time (s).
Unit Prefixes
Once we have defined the fundamental units, it is easy to introduce larger and
smaller units for the same physical quantities. In the metric system these other
units are related to the fundamental units (or, in the case of mass, to the gram) by
1
multiples of 10 or 10
Thus one kilometer 11 km2 is 1000 meters, and one centi1
1
meter 11 cm2 is 100 meter. We usually express multiples of 10 or 10
in exponential
3 1
-3
notation: 1000 = 10 , 1000 = 10 , and so on. With this notation, 1 km = 103 m
and 1 cm = 10-2 m.
The names of the additional units are derived by adding a prefix to the name
of the fundamental unit. For example, the prefix “kilo-,” abbreviated k, always
means a unit larger by a factor of 1000; thus
1 kilometer = 1 km = 103 meters = 103 m
1 kilogram = 1 kg = 103 grams = 103 g
1 kilowatt
= 1 kW = 103 watts
= 103 W
A table in Appendix A lists the standard SI units, with their meanings and
abbreviations.
Table 1.1 gives some examples of the use of multiples of 10 and their prefixes
with the units of length, mass, and time. Figure 1.5 (next page) shows how these
prefixes are used to describe both large and small distances.
Table 1.1 Some Units of Length, Mass, and Time
Length
Mass
-9
Time
1 nanometer = 1 nm = 10 m
(a few times the size of the largest atom)
1 microgram = 1 mg = 10 g = 10 kg
(mass of a very small dust particle)
1 nanosecond = 1 ns = 10-9 s
(time for light to travel 0.3 m)
1 micrometer = 1 mm = 10-6 m
(size of some bacteria and other cells)
1 milligram = 1 mg = 10-3 g = 10-6 kg
(mass of a grain of salt)
1 microsecond = 1 ms = 10-6 s
(time for space station to move 8 mm)
1 millimeter = 1 mm = 10-3 m
(diameter of the point of a ballpoint pen)
1 gram
= 1 g = 10-3 kg
(mass of a paper clip)
1 millisecond = 1 ms = 10-3 s
(time for a car moving at freeway speed
to travel 3 cm)
1 centimeter = 1 cm = 10-2 m
(diameter of your little finger)
1 kilometer = 1 km = 103 m
(distance in a 10-minute walk)
-6
-9
5
6
Chapter 1 Units, physical Quantities, and Vectors
1.5 Some typical lengths in the universe.
Note: (f) is a scanning tunneling
microscope image of atoms on a
crystal surface; (g) is an artist’s
impression.
(a) 10 26 m
Limit of the
observable
universe
(b) 1011 m
Distance to
the sun
(c) 107 m
Diameter of
the earth
1.6 Many everyday items make use of
both SI and British units. An example is
this speedometer from a U.S.-built automobile, which shows the speed in both
kilometers per hour (inner scale) and miles
per hour (outer scale).
(d) 1 m
Human
dimensions
(e) 10 -5 m
Diameter of a
red blood cell
(f) 10 -10 m
Radius of an
atom
(g) 10 -14 m
Radius of an
atomic nucleus
The British System
Finally, we mention the British system of units. These units are used in only the
United States and a few other countries, and in most of these they are being replaced
by SI units. British units are now officially defined in terms of SI units, as follows:
Length:
Force:
1 inch = 2.54 cm (exactly)
1 pound = 4.448221615260 newtons (exactly)
The newton, abbreviated N, is the SI unit of force. The British unit of time is the
second, defined the same way as in SI. In physics, British units are used in mechanics and thermodynamics only; there is no British system of electrical units.
In this book we use SI units for all examples and problems, but we occasionally give approximate equivalents in British units. As you do problems using
SI units, you may also wish to convert to the approximate British equivalents if
they are more familiar to you (Fig. 1.6). But you should try to think in SI units as
much as you can.
1.4 USing and ConverTing UniTS
CaUTion Always use units in calcula­
tions Make it a habit to always write
numbers with the correct units and carry
the units through the calculation as in the
example above. This provides a very useful check. If at some stage in a calculation
you find that an equation or an expression
has inconsistent units, you know you have
made an error. In this book we will always
carry units through all calculations, and
we strongly urge you to follow this practice
when you solve problems. ❙
We use equations to express relationships among physical quantities, represented
by algebraic symbols. Each algebraic symbol always denotes both a number and
a unit. For example, d might represent a distance of 10 m, t a time of 5 s, and v a
speed of 2 m>s.
An equation must always be dimensionally consistent. You can’t add apples
and automobiles; two terms may be added or equated only if they have the same
units. For example, if a body moving with constant speed v travels a distance d in
a time t, these quantities are related by the equation
d = vt
If d is measured in meters, then the product vt must also be expressed in meters.
Using the above numbers as an example, we may write
10 m = a 2
m
b (5 s)
s
Because the unit s in the denominator of m>s cancels, the product has units of
meters, as it must. In calculations, units are treated just like algebraic symbols
with respect to multiplication and division.
1.4 Using and Converting Units
Problem-Solving STraTegy 1.2
7
Solving PhySicS ProblemS
idenTify the relevant concepts: In most cases, it’s best to use the
fundamental SI units (lengths in meters, masses in kilograms, and
times in seconds) in every problem. If you need the answer to be
in a different set of units (such as kilometers, grams, or hours),
wait until the end of the problem to make the conversion.
seT uP the problem and execuTe the solution: Units are multiplied and divided just like ordinary algebraic symbols. This gives
us an easy way to convert a quantity from one set of units to another: Express the same physical quantity in two different units
and form an equality.
For example, when we say that 1 min = 60 s, we don’t mean
that the number 1 is equal to the number 60; rather, we mean that
1 min represents the same physical time interval as 60 s. For this
reason, the ratio (1 min)>(60 s) equals 1, as does its reciprocal,
(60 s)>(1 min). We may multiply a quantity by either of these
factors (which we call unit multipliers) without changing that
quantity’s physical meaning. For example, to find the number of
seconds in 3 min, we write
3 min = (3 min) a
60 s
b = 180 s
1 min
evaLuaTe your answer: If you do your unit conversions correctly,
unwanted units will cancel, as in the example above. If, instead,
you had multiplied 3 min by (1 min)>(60 s), your result would
1
have been the nonsensical 20
min2>s. To be sure you convert units
properly, include the units at all stages of the calculation.
Finally, check whether your answer is reasonable. For example, the result 3 min = 180 s is reasonable because the second is
a smaller unit than the minute, so there are more seconds than
minutes in the same time interval.
Solution
examPle 1.1 converTing SPeed uniTS
The world land speed record of 763.0 mi>h was set on October 15,
1997, by Andy Green in the jet-engine car Thrust SSC. Express
this speed in meters per second.
763.0 mi>h = a763.0
mi 1.609 km 1000 m
1h
b
ba
ba
ba
h
1 mi
1 km
3600 s
= 341.0 m>s
soLuTion
idenTify, seT uP, and execuTe: We need to convert the units
of a speed from mi>h to m>s. We must therefore find unit multipliers that relate (i) miles to meters and (ii) hours to seconds. In
Appendix E we find the equalities 1 mi = 1.609 km, 1 km =
1000 m, and 1 h = 3600 s. We set up the conversion as follows,
which ensures that all the desired cancellations by division take place:
evaLuaTe: This example shows a useful rule of thumb: A speed
expressed in m>s is a bit less than half the value expressed in mi>h, and
a bit less than one-third the value expressed in km>h. For example, a
normal freeway speed is about 30 m>s = 67 mi>h = 108 km>h, and
a typical walking speed is about 1.4 m>s = 3.1 mi>h = 5.0 km>h.
Solution
examPle 1.2 converTing volume uniTS
One of the world’s largest cut diamonds is the First Star of Africa
(mounted in the British Royal Sceptre and kept in the Tower of
London). Its volume is 1.84 cubic inches. What is its volume in
cubic centimeters? In cubic meters?
soLuTion
idenTify, seT uP, and execuTe: Here we are to convert the units
of a volume from cubic inches 1in.32 to both cubic centimeters
1cm32 and cubic meters 1m32. Appendix E gives us the equality
1 in. = 2.540 cm, from which we obtain 1 in.3 = (2.54 cm)3. We
then have
1.84 in.3 = 11.84 in.32a
2.54 cm 3
b
1 in.
= 11.84212.5423
in.3 cm3
in.3
= 30.2 cm3
Appendix E also gives us 1 m = 100 cm, so
30.2 cm3 = 130.2 cm32a
= 130.22a
1m 3
b
100 cm
1 3 cm3 m3
b
= 30.2 * 10-6 m3
100
cm3
= 3.02 * 10-5 m3
evaLuaTe: Following the pattern of these conversions, can you
show that 1 in.3 ≈ 16 cm3 and that 1 m3 ≈ 60,000 in.3?
8
ChaPter 1 Units, Physical Quantities, and Vectors
1.7 This spectacular mishap was the
result of a very small percent error—
traveling a few meters too far at the end of
a journey of hundreds of thousands of
meters.
Table 1.2 using significant figures
Multiplication or division:
Result can have no more significant figures
than the factor with the fewest significant figures:
0.745 * 2.2
= 0.42
3.885
1.32578 * 107 * 4.11 * 10 - 3 = 5.45 * 104
Addition or subtraction:
Number of significant figures is determined by
the term with the largest uncertainty (i.e., fewest
digits to the right of the decimal point):
27.153 + 138.2 - 11.74 = 153.6
1.8 Determining the value of p from the
circumference and diameter of a circle.
135 mm
424 mm
The measured values have only three significant
figures, so their calculated ratio (p) also has
only three significant figures.
1.5 uncerTainTy and significanT figures
Measurements always have uncertainties. If you measure the thickness of the cover
of a hardbound version of this book using an ordinary ruler, your measurement is
reliable to only the nearest millimeter, and your result will be 3 mm. It would be
wrong to state this result as 3.00 mm; given the limitations of the measuring device,
you can’t tell whether the actual thickness is 3.00 mm, 2.85 mm, or 3.11 mm. But
if you use a micrometer caliper, a device that measures distances reliably to the
nearest 0.01 mm, the result will be 2.91 mm. The distinction between the measurements with a ruler and with a caliper is in their uncertainty; the measurement
with a caliper has a smaller uncertainty. The uncertainty is also called the error
because it indicates the maximum difference there is likely to be between the measured value and the true value. The uncertainty or error of a measured value depends on the measurement technique used.
We often indicate the accuracy of a measured value—that is, how close it is
likely to be to the true value—by writing the number, the symbol {, and a second number indicating the uncertainty of the measurement. If the diameter of a
steel rod is given as 56.47 { 0.02 mm, this means that the true value is likely to
be within the range from 56.45 mm to 56.49 mm. In a commonly used shorthand
notation, the number 1.64541212 means 1.6454 { 0.0021. The numbers in parentheses show the uncertainty in the final digits of the main number.
We can also express accuracy in terms of the maximum likely fractional
error or percent error (also called fractional uncertainty and percent uncertainty). A resistor labeled ;47 ohms { 10%< probably has a true resistance
that differs from 47 ohms by no more than 10% of 47 ohms—that is, by about
5 ohms. The resistance is probably between 42 and 52 ohms. For the diameter
of the steel rod given above, the fractional error is 10.02 mm2>156.47 mm2, or
about 0.0004; the percent error is 10.000421100%2, or about 0.04%. Even small
percent errors can be very significant (Fig. 1.7).
In many cases the uncertainty of a number is not stated explicitly. Instead, the uncertainty is indicated by the number of meaningful digits, or significant figures, in
the measured value. We gave the thickness of the cover of the book as 2.91 mm, which
has three significant figures. By this we mean that the first two digits are known to
be correct, while the third digit is uncertain. The last digit is in the hundredths place,
so the uncertainty is about 0.01 mm. Two values with the same number of significant
figures may have different uncertainties; a distance given as 137 km also has three
significant figures, but the uncertainty is about 1 km. A distance given as 0.25 km
has two significant figures (the zero to the left of the decimal point doesn’t count); if
given as 0.250 km, it has three significant figures.
When you use numbers that have uncertainties to compute other numbers, the
computed numbers are also uncertain. When numbers are multiplied or divided,
the result can have no more significant figures than the factor with the fewest significant figures has. For example, 3.1416 * 2.34 * 0.58 = 4.3. When we add
and subtract numbers, it’s the location of the decimal point that matters, not the
number of significant figures. For example, 123.62 + 8.9 = 132.5. Although
123.62 has an uncertainty of about 0.01, 8.9 has an uncertainty of about 0.1. So
their sum has an uncertainty of about 0.1 and should be written as 132.5, not
132.52. Table 1.2 summarizes these rules for significant figures.
To apply these ideas, suppose you want to verify the value of p, the ratio of
the circumference of a circle to its diameter. The true value of this ratio to ten
digits is 3.141592654. To test this, you draw a large circle and measure its circumference and diameter to the nearest millimeter, obtaining the values 424 mm
and 135 mm (Fig. 1.8). You punch these into your calculator and obtain the quotient 1424 mm2>1135 mm2 = 3.140740741. This may seem to disagree with the
true value of p, but keep in mind that each of your measurements has three significant figures, so your measured value of p can have only three significant
figures. It should be stated simply as 3.14. Within the limit of three significant
figures, your value does agree with the true value.
1.5 Uncertainty and Significant Figures
9
In the examples and problems in this book we usually give numerical values with three significant figures, so your answers should usually have no more
than three significant figures. (Many numbers in the real world have even less
accuracy. An automobile speedometer, for example, usually gives only two significant figures.) Even if you do the arithmetic with a calculator that displays ten
digits, a ten-digit answer would misrepresent the accuracy of the results. Always
round your final answer to keep only the correct number of significant figures or,
in doubtful cases, one more at most. In Example 1.1 it would have been wrong to
state the answer as 341.01861 m>s. Note that when you reduce such an answer
to the appropriate number of significant figures, you must round, not truncate.
Your calculator will tell you that the ratio of 525 m to 311 m is 1.688102894; to
three significant figures, this is 1.69, not 1.68.
When we work with very large or very small numbers, we can show significant figures much more easily by using scientific notation, sometimes called
powers-of-10 notation. The distance from the earth to the moon is about
384,000,000 m, but writing the number in this form doesn’t indicate the number
of significant figures. Instead, we move the decimal point eight places to the left
(corresponding to dividing by 108) and multiply by 108; that is,
384,000,000 m = 3.84 * 108 m
In this form, it is clear that we have three significant figures. The number
4.00 * 10-7 also has three significant figures, even though two of them are
zeros. Note that in scientific notation the usual practice is to express the quantity
as a number between 1 and 10 multiplied by the appropriate power of 10.
When an integer or a fraction occurs in an algebraic equation, we treat
that number as having no uncertainty at all. For example, in the equation
v x2 = v 0x 2 + 2ax 1x - x02, which is Eq. (2.13) in Chapter 2, the coefficient 2 is
exactly 2. We can consider this coefficient as having an infinite number of significant figures (2.000000 c). The same is true of the exponent 2 in v x2 and v 0x 2.
Finally, let’s note that precision is not the same as accuracy. A cheap digital
watch that gives the time as 10:35:17 a.m. is very precise (the time is given to
the second), but if the watch runs several minutes slow, then this value isn’t very
accurate. On the other hand, a grandfather clock might be very accurate (that is,
display the correct time), but if the clock has no second hand, it isn’t very precise.
A high-quality measurement is both precise and accurate.
Solution
examPle 1.3 SignificanT figureS in mulTiPlicaTion
The rest energy E of an object with rest mass m is given by Albert
Einstein’s famous equation E = mc2, where c is the speed of light
in vacuum. Find E for an electron for which (to three significant
figures) m = 9.11 * 10-31 kg. The SI unit for E is the joule (J);
1 J = 1 kg # m2>s2.
soLuTion
idenTify and seT uP: Our target variable is the energy E. We are
given the value of the mass m; from Section 1.3 (or Appendix F) the
speed of light is c = 2.99792458 * 108 m>s.
execuTe: Substituting the values of m and c into Einstein’s equa-
tion, we find
E = 19.11 * 10-31 kg212.99792458 * 108 m>s22
= 19.11212.9979245822110-312110822 kg # m2>s2
= 181.8765967821103 - 31 + 12 * 8242 kg # m2>s2
= 8.187659678 * 10-14 kg # m2>s2
Since the value of m was given to only three significant figures,
we must round this to
E = 8.19 * 10-14 kg # m2>s2 = 8.19 * 10-14 J
evaLuaTe: While the rest energy contained in an electron may seem
ridiculously small, on the atomic scale it is tremendous. Compare
our answer to 10-19 J, the energy gained or lost by a single atom
during a typical chemical reaction. The rest energy of an electron is
about 1,000,000 times larger! (We’ll discuss the significance of rest
energy in Chapter 37.)
10
ChaPter 1 Units, Physical Quantities, and Vectors
TesT Your undersTanding of secTion 1.5 The density of a material
is equal to its mass divided by its volume. What is the density 1in kg>m32 of a rock of
mass 1.80 kg and volume 6.0 * 10-4 m3? (i) 3 * 103 kg>m3; (ii) 3.0 * 103 kg >m3;
(iii) 3.00 * 103 kg>m3; (iv) 3.000 * 103 kg>m3; (v) any of these—all of these answers
are mathematically equivalent. ❙
1.6 esTimaTes and orders of magniTude
PhET: Estimation
We have stressed the importance of knowing the accuracy of numbers that represent physical quantities. But even a very crude estimate of a quantity often gives
us useful information. Sometimes we know how to calculate a certain quantity,
but we have to guess at the data we need for the calculation. Or the calculation
might be too complicated to carry out exactly, so we make rough approximations. In either case our result is also a guess, but such a guess can be useful even
if it is uncertain by a factor of two, ten, or more. Such calculations are called
order-of-magnitude estimates. The great Italian-American nuclear physicist
Enrico Fermi (1901–1954) called them “back-of-the-envelope calculations.”
Exercises 1.17 through 1.23 at the end of this chapter are of the estimating, or
order-of-magnitude, variety. Most require guesswork for the needed input data.
Don’t try to look up a lot of data; make the best guesses you can. Even when they
are off by a factor of ten, the results can be useful and interesting.
Solution
ExamplE 1.4 an ordEr-of-magnitudE EstimatE
You are writing an adventure novel in which the hero escapes with
a billion dollars’ worth of gold in his suitcase. Could anyone carry
that much gold? Would it fit in a suitcase?
soLuTion
idenTifY, seT up, and execuTe: Gold sells for about $1400 an
1
ounce, or about $100 for 14
ounce. (The price per ounce has
varied between $200 and $1900 over the past twenty years or
so.) An ounce is about 30 grams, so $100 worth of gold has a
1
mass of about 14
of 30 grams, or roughly 2 grams. A billion
9
110 2 dollars’ worth of gold has a mass 107 times greater, about
application scalar Temperature,
vector Wind The comfort level on a win-
try day depends on the temperature, a scalar
quantity that can be positive or negative (say,
+ 5 °C or - 2 0 °C ) but has no direction. It also
depends on the wind velocity, a vector quantity with both magnitude and direction (for
example, 15 km>h from the west).
2 * 107 120 million2 grams or 2 * 104 120,0002 kilograms. A
thousand kilograms has a weight in British units of about a ton,
so the suitcase weighs roughly 20 tons! No human could lift it.
Roughly what is the volume of this gold? The density of water is
103 kg>m3; if gold, which is much denser than water, has a density
10 times greater, then 104 kg of gold fit into a volume of 1 m3. So
109 dollars’ worth of gold has a volume of 2 m3, many times the
volume of a suitcase.
evaLuaTe: Clearly your novel needs rewriting. Try the calculation again with a suitcase full of five-carat (1-gram) diamonds, each
worth $500,000. Would this work?
TesT Your undersTanding of secTion 1.6 Can you estimate the total
number of teeth in the mouths of all the students on your campus? (Hint: How many
teeth are in your mouth? Count them!) ❙
1.7 vecTors and vecTor addiTion
Some physical quantities, such as time, temperature, mass, and density, can be
described completely by a single number with a unit. But many other important
quantities in physics have a direction associated with them and cannot be described by a single number. A simple example is the motion of an airplane: We
must say not only how fast the plane is moving but also in what direction. The
speed of the airplane combined with its direction of motion constitute a quantity
called velocity. Another example is force, which in physics means a push or pull
exerted on a body. Giving a complete description of a force means describing
both how hard the force pushes or pulls on the body and the direction of the push
or pull.
11
1.7 Vectors and Vector addition
When a physical quantity is described by a single number, we call it a
scalar quantity. In contrast, a vector quantity has both a magnitude (the
“how much” or “how big” part) and a direction in space. Calculations that combine scalar quantities use the operations of ordinary arithmetic. For example,
6 kg + 3 kg = 9 kg, or 4 * 2 s = 8 s. However, combining vectors requires a
different set of operations.
To understand more about vectors and how they combine, we start with the
simplest vector quantity, displacement. Displacement is a change in the position
of an object. Displacement is a vector quantity because we must state not only
how far the object moves but also in what direction. Walking 3 km north from
your front door doesn’t get you to the same place as walking 3 km southeast;
these two displacements have the same magnitude but different directions.
We usually
represent a vector quantity such as displacement by a single letter,
S
such as A in Fig. 1.9a. In this book we always print vector symbols in boldface
italic type with an arrow above them. We do this to remind you that vector quantities have different properties from scalar quantities; the arrow is a reminder that
vectors have direction. When you handwrite a symbol for a vector, always write
it with an arrow on top. If you don’t distinguish between scalar and vector quantities in your notation, you probably won’t make the distinction in your thinking
either, and confusion will result.
We always draw a vector as a line with an arrowhead at its tip. The length of
the line shows the vector’s magnitude, and the direction of the arrowhead shows
the vector’s direction. Displacement is always a straight-line segment directed
from the starting point to the ending point, even though the object’s actual path
may be curved (Fig. 1.9b). Note that displacement is not related directly to the
total distance traveled. If the object were to continue past P2 and then return to
P1 , the displacement for the entire trip would be zero (Fig. 1.9c).
If two vectors have the same direction, they are parallel. If they have the same
magnitude and the same
direction, they are equal, no matter where they are located
S
in space. The vector A ′ fromS point P3 to point P4 in Fig. 1.10 has the same length
and direction as the vector A from P1 to P2 . These two displacements
are equal,
S
S
even though they start at different points. We write this as A ′ = A in Fig. 1.10;
the boldface equals sign emphasizes that equality of two vector quantities is not
the same relationship as equality of two scalar quantities. Two vector quantities are
equal only when
they have the same magnitude and theSsame direction.
S
Vector B in Fig.
1.10, however, is not equal to A because its direction is
S
opposite that of A. We define the negative of a vector as a vector having the
same magnitude as Sthe original vector Sbut the opposite direction. The negative
of vector quantity A is denoted as −A, and we Suse a boldface minus signS to
emphasize the vector nature of the quantities. If A is 87S m south,
then −A is
S
87
m
north.
Thus
we
can
write
the
relationship
between
A
and
B
in
Fig. 1.10 as
S
S
S
S
S
S
A = − B or B = − A. When two vectors A and B have opposite directions,
whether their magnitudes are the same or not, we say that they are antiparallel.
We usually represent the magnitude of a vector quantity by the same letter
used for the vector, but in
lightface italic type with no arrow on top. For example,
S
if displacement vector A is 87 m south, then A = 87 m. An alternative notation
is the vector symbol with vertical bars on both sides:
S
S
1Magnitude of A2 = A = 0 A 0
1.9 Displacement as a vector quantity.
(a) We represent a displacement by an arrow that
points in the direction of displacement.
Ending position: P2
S
Displacement A
Starting position: P1
Handwritten notation:
(b) A displacement is always a straight arrow
directed from the starting position to the ending
position. It does not depend on the path taken,
even if the path is curved.
P2
S
A
Path taken
P1
(c) Total displacement for a round trip is 0,
regardless of the path taken or distance traveled.
P1
1.10 The meaning of vectors that have the
same magnitude and the same or opposite
direction.
P2
P4
P5
(1.1)
The magnitude of a vector quantity is a scalar quantity (a number) and is always
positive. Note that a vector can never be equalS to a scalar because they are
different kinds of quantities. The expression ;A = 6 m< is just as wrong as
;2 oranges = 3 apples<!
When we draw diagrams with vectors, it’s best to use a scale similar to those
used for maps. For example, a displacement of 5 km might be represented in a diagram by a vector 1 cm long, and a displacement of 10 km by a vector 2 cm long.
S
S
P1
S
S
A′ = A
A
S
B = −A
P3
P6
S
S
Displacements A and A′
are equal because they
have the same length
and direction.
S
Displacement B has
theSsame magnitude
as A but opposite
S
direction; BSis the
negative of A.
12
ChaPter 1 Units, Physical Quantities, and Vectors
1.11 Three ways to add two vectors.
vector addition and subtraction
(a) We can add two vectors by placing them
head to tail.
Suppose
a particle undergoes a displacement A followed by a second displaceS
ment B. The final result is the same as if the particle
had started at the same
S
initial pointS and undergone a single displacement C (Fig. 1.11a).
WeS call disS
placement C the vector sum, or resultant, of displacements A and B. We express this relationship symbolically as
S
The vector sum C
extends from Sthe
tail of vector A ...
... to the head
S
of vector B.
S
B
S
A
S
S
S
S
S
(1.2)
The boldface plus sign emphasizes that adding two vector quantities requires a
geometrical process and is not the same operation as adding two scalar quantities such as 2 + 3 = 5. In vector addition we usually place the tail of the second
vector at the head, or tip, of the first
vectorS (Fig. 1.11a).
S
S
S
If we make the displacements A and B in reverse order, with B first and A
second, the result is the same (Fig. 1.11b). Thus
(b) Adding
them
in
reverse
order gives the same
S
S
S
S
result: A + B = B + A. The order doesn’t
matter in vector addition.
S
S
C=A+B
S
C=A+ B
C=B+ A
S
S
S
A
S
S
S
S
S
S
S
C = B + A and A + B = B + A
S
B
This shows that the order of terms in a vector sum doesn’t matter. In other words,
vector addition obeys the commutative law.
Figure
1.11c
shows another way to represent the vector
sum: If we draw vecS
S
S
tors A and B with their tailsSat theSsame point, vector C is the diagonal of a parallelogram constructed with A and B as two adjacent sides.
(c) We can also add two vectors by placing them
tail to tail and constructing a parallelogram.
S
A
S
S
(1.3)
S
C=A+ B
S
B
S
S
S
(a) Only when vectors A and B are parallel
S
does the magnitude of their vector sum C equal
the sum of their magnitudes: C = A + B.
S
S
A
B
S
S
S
S
(b) When A and B are antiparallel,
the
S
magnitude of their vector sum C equals the
difference of their magnitudes: C = 0 A - B 0.
S
S
S
S
S
S
S
S
S
S
S
S
Alternatively,
we can first
add B and C to obtain vector E (Fig. 1.13c), and then
S
S
S
add A and E to obtain R:
S
S
S
R = 1A + B2 + C = D + C
A
C=A+ B
S
Figure 1.13a shows three vectors
A, SB, and C. To find the vector
sum of all
S
S
three, inS Fig. 1.13b
we
first
add
A
and
B
to
give
a
vector
sum
D
;
we
then add
S
S
vectors C and D by the same process to obtain the vector sum R:
S
C=A+ B
S
S
then magnitude C equals magnitude A plus magnitude B. In general, this
conclusion
is
S
S
wrong; for the vectors shown
in
Fig.
1.11,
C
6
A
+
B.
The
magnitude
of
A
+
B
depends
S
S
S
S
on the magnitudes
of SA and B and on the angle between
A and
BS. Only in the special
S
S
S
case in which A and
B areS parallel is the magnitude of C = A + B equal to the sum of
S
the magnitudes of SA and B (Fig. 1.12a). When the vectors are antiparallel
(Fig. 1.12b),
S
S
the magnitude of C equals the difference of the magnitudes of A and B. Be careful to
distinguish between scalar and vector quantities, and you’ll avoid making errors about the
magnitude of a vector sum. ❙
1.12 Adding vectors that are (a) parallel
and (b) antiparallel.
S
S
cauTion magnitudes in vector addition It’s a common error to conclude that if C = A + B,
S
S
S
S
S
R = A + 1B + C 2 = A + E
B
S
S
S
S
We don’t
even need to draw vectors D and E; all we need to do is draw A, B,
S
and C in succession,
with the tail of each at the head of the one preceding it. The
S
sum vector R extends from the tail of the first vector to the head of the last vector
S
S
S
1.13 Several constructions for finding the vector sum A + B + C.
(a) To find the sum of
these three vectors ...
S
S
D
B
S
B
S
C
E
A
S
B
S
S
(e)
... or add A, B, and
S
C in anySother order and
still get R.
S
R
S
S
S
S
(d)
... or Sadd A, B, and
S
C to get R directly ...
R
S
A
S
S
R
S
S
S
(c) ... or
add B and C
S
to
get
E
and
then
add
S
S
S
A to E to get R ...
S
S
C
A
S
(b) ... add
A and B
S
to
get
D
and
then add
S
S
C to D to get theSfinal
sum (resultant) R ...
S
R
S
S
C
A
S
C
S
B
S
B
S
A
S
C
1.7 Vectors and Vector addition
S
S
S
13
S
1.14 To construct
the vector
difference A − B, you can either place the tail of −B at the head of A or place
S
S
the two vectors A and B head to head.
S
S
S
S
A
−
S
B
S
S
... is equivalent to adding −B to A.
Subtracting B from A ...
S
A
=
+
S
−B
=
S
S
A
S
S
A + 1−B2
S
S
=A−B
S
−B
S
S
=
S
S
S
S
S
A − B = A + 1−B2
S
S
B
S
With
A and B head to head,
S
S
A − BSis the vector from
the
S
tail of A to the tail of B.
PhET: Vector Addition
(1.4)
Figure 1.14 shows an example of vector subtraction.
A vector quantity such as a displacement can
be multiplied by a scalar quanS
tity (an ordinary number). The displacement
2A
is
a displacement (vector quanS
tity) in the
same
direction
as
vector
A
but
twice
as long; this is the same
as
S
S
adding A to itself (Fig.
1.15a).
In
general,
when
we
multiply
a
vector
A
by
a
S
scalar c, the result cA hasS magnitude 0 c 0 A (the
absolute
value
of
c
multiplied
by
S
S
the magnitudeS of vector A). If c is positive, cASis in the same
direction asS A; if c
S
is negative,
cA is in theS direction opposite to A. Thus 3A is parallel to A, while
S
-3A is antiparallel to A (Fig. 1.15b).
A scalar used to multiply a vector can also Sbe a physical quantity. SFor examS
ple, you may be familiar with the relationship F = ma; the net force F (a vector
quantity) that acts on a body is equal to the product of the body’s mass
m (a scalar
S
S
quantity) and its acceleration a (a vector quantity). The direction
of
F
is
the same
S
S
as that of a because m is positive, and the magnitude of F is equal to the mass m
S
multiplied by the magnitude of a. The unit of force is the unit of mass multiplied
by the unit of acceleration.
1.15 Multiplying a vector by a scalar.
(a) Multiplying a vector by a positive scalar
changes the magnitude (length) of the vector
but not its direction.
S
A
S
2A
S
S
2A is twice as long as A.
(b) Multiplying a vector by a negative scalar
changes its magnitude and reverses its direction.
S
A
S
- 3A
S
S
- 3A is three times as long as A and points
in the opposite direction.
Solution
examPle 1.5 adding Two vecTorS aT righT angleS
A cross-country skier skis 1.00 km north and then 2.00 km east on
a horizontal snowfield. How far and in what direction is she from
the starting point?
1.16 The vector diagram, drawn to scale, for a ski trip.
N
W
soLuTion
E
S
2.00 km
idenTify and seT uP: The problem involves combining two
displacements at right angles to each other. This vector addition amounts to solving a right triangle, so we can use the
Pythagorean theorem and simple trigonometry. The target variables are the skier’s straight-line distance and direction from
her starting point. Figure 1.16 is a scale diagram of the two
displacements and the resultant net displacement. We denote
the direction from the starting point by the angle f (the Greek
letter phi). The displacement appears to be a bit more than
2 km. Measuring the angle with a protractor indicates that f is
about 63°.
S
A−B
S
A
S
With
A and −B head to tail,
S
S
A − BSis the vector from Sthe
tail of A to the head of −B.
(Fig. 1.13d). The order makes no difference; Fig. 1.13e shows a different order,
and you should try others. Vector addition obeys the associative law.
We can subtract vectors as well
as add them. To see how, recall that vector
S
S
−A hasSthe same
magnitude
as
A
but
the opposite direction. We
defineSthe difS
S
S
S
ference A − B of two vectors A and B to be the vector sum of A and −B:
S
A + 1−B2 = A − B
1.00 km
0
f
Resultant displacement
1 km
2 km
Continued
14
ChaPter 1 Units, Physical Quantities, and Vectors
execuTe: The distance from the starting point to the ending point is
equal to the length of the hypotenuse:
We can describe the direction as 63.4° east of north or
90° - 63.4° = 26.6° north of east.
211.00 km22 + 12.00 km22 = 2.24 km
evaLuaTe: Our answers (2.24 km and f = 63.4°) are close to
A little trigonometry (from Appendix B) allows us to find angle f:
tan f =
our predictions. In Section 1.8 we’ll learn how to easily add two
vectors not at right angles to each other.
Opposite side
2.00 km
=
= 2.00
Adjacent side
1.00 km
f = arctan 2.00 = 63.4°
daTa SpeakS
vector addition and subtraction
When students were given a problem
about adding or subtracting two vectors,
more than 28% gave an incorrect answer.
Common errors:
●
●
When
adding
vectors,
drawing vectors
S S
S
S
A, B, and A + B incorrectly. The headto-tail arrangement shown in Figs. 1.11a
and 1.11b is easiest.
When subtracting
vectors,
drawing
S S
S
S
vectors A, B, and A − B incorrectly.
S
S
Remember that subtracting
B from A is
S
S
the same as adding − B to A (Fig. 1.14).
S
1.17 Representing a vector A in terms of
its components Ax and Ay.
S
y
The components of A
are the projections
of the vector onto
the x- and y-axes.
A
Ax = Acos u
In this case, both Ax and Ay are positive.
In Section 1.7 we added vectors by using a scale diagram and properties of right
triangles. Making measurements of a diagram offers only very limited accuracy,
and calculations with right triangles work only when the two vectors are perpendicular. So we need a simple but general method for adding vectors. This is
called the method of components.
S
To define what we mean by the components of a vector A, we begin with Sa rectangular (Cartesian) coordinate system
of axes (Fig. 1.17). If we think of A as a
S
displacement vector, we can regard A as the sum of a displacement parallel to the
x-axis and a displacement parallel to the y-axis. We use the numbers Ax and Ay to
tell us how much displacement there is parallel to the x-axis and how much there is
parallel to the y-axis, respectively.
For example, if the +x-axis points east and the
S
+y-axis points north, A in Figure 1.17 could be the sum of a 2.00-m displacement
to the east and a 1.00-m displacement to the north. Then Ax = +2.00 m and
Ay = +1.00 m. We can use the same idea for any vectors, not justS displacement
vectors. The two numbers Ax and Ay are called the components of A.
S
Ay = Asinu
u
1.8 comPonenTs of vecTors
cauTion components are not vectors The components Ax and Ay of a vector A are
S
O
TesT
your
undersTanding of secTion 1.7 Two displacement vectors,
S
S
S and T, have magnitudes S = 3 m Sand TS = 4 m. Which of the following could be the
magnitude of the difference vector S − T? (There may be more than one correct answer.) (i) 9 m; (ii) 7 m; (iii) 5 m; (iv) 1 m; (v) 0 m; (vi) -1 m. ❙
x
numbers; they are not vectors themselves. This is why we print the symbols for components in lightface italic type with no arrow on top instead of in boldface italic with an
arrow, which is reserved for vectors. ❙
S
?
We can calculate the components of vector A if we know its magnitude A
and its direction. We’ll describe the direction of a vector by its angle relative
to some reference direction. In Fig. 1.17
this reference direction is the positive
S
x-axis, and the angle between vector
A
and
the positive x-axis is u (the Greek
S
letter theta). Imagine that vector A originally lies along the +x@axis and that
you then rotate it to its true direction, as indicated by the arrow in Fig. 1.17 on
the arc for angle u. If this rotation is from the +x@axis toward the +y@axis, as
is the case in Fig. 1.17, then u is positive; if the rotation is from the +x@axis
toward the -y@axis, then u is negative. Thus the +y@axis is at an angle of 90°, the
-x@axis at 180°, and the -y@axis at 270° (or -90°). If u is measured in this way,
then from the definition of the trigonometric functions,
Ay
Ax
= cos u
and
= sin u
A
A
(1.5)
Ax = A cos u
and
Ay = A sin u
1u measured from the +x@axis, rotating toward the +y@axis2
1.8 Components of Vectors
In Fig. 1.17 Ax and Ay are positive. This is consistent with Eqs. (1.5); u is in the
first quadrant (between 0° and 90°), and both the cosine and the sine of an angle
in this quadrant are positive. But in Fig. 1.18a the component Bx is negative and
the component
By is positive. (If the +x-axis points east and the +y-axis points
S
north, B could represent a displacement of 2.00 m west and 1.00 m north. Since
west is in the –x-direction and north is in the +y-direction, Bx = -2.00 m is
negative and By = +1.00 m is positive.) Again, this is consistent with Eqs. (1.5);
now u is in the second quadrant, so cos u is negative and sin u is positive. In
Fig. 1.18b both Cx and Cy are negative (both cos u and sin u are negative in the
third quadrant).
1.18 The components of a vector may be
positive or negative numbers.
(a)
y
are correct only when the angle u is measured from the positive x-axis. If the angle of the
vector is given from a different reference direction or you use a different rotation direction, the relationships are different! Example 1.6 illustrates this point. ❙
By is positive.
S
B
By 1+2
u
Bx 1-2
x
Bx is negative.
(b)
cauTion relating a vector’s magnitude and direction to its components Equations (1.5)
15
y
u
Cx 1-2
x
Cy 1-2
S
C
S
Both components of C are negative.
Solution
examPle 1.6 finding comPonenTS
S
(a) What are the x- and y-components of vector D in Fig. 1.19a?
The magnitude of the vector is D = 3.00 m, and angle
S
a = 45°. (b) What are the x- and y-components of vector E
in Fig. 1.19b? The magnitude of the vector is E = 4.50 m, and
angle b = 37.0°.
components in part (a) are roughly 2 m, and that those in part (b)
are 3 m and 4 m. The figure indicates the signs of the components.
execuTe: (a) The angle a (the Greek letter alpha) between the
S
positive x-axis and D is measured toward the negative y-axis.
The angle we must use in Eqs. (1.5) is u = -a = - 45°. We
then find
soLuTion
Dx = D cos u = 13.00 m21cos1-45°22 = + 2.1 m
idenTify and seT uP: We can use Eqs. (1.5) to find the compo-
nents of these vectors, but we must be careful: Neither angle a
nor b in Fig. 1.19 is measured from the +x@axis toward the
+ y@axis. We estimate from the figure that the lengths of both
1.19 Calculating the x- and y-components of vectors.
(a)
Angle a is
measured in the
y wrong sense from
the + x-axis, so in
Eqs. (1.5) we
must use -a.
Dx 1+2
a
Dy 1-2
(b)
Ey 1+2
x
S
D
Angle b is measured from the
+ y-axis, not from the +x-axis.
y
u
b
Ex 1+2
S
E
x
We must use u,
which is measured from
the + x-axis toward the
+ y-axis, in Eqs. (1.5).
Dy = D sin u = 13.00 m21sin1-45°22 = - 2.1 m
Had we carelessly substituted +45° for u in Eqs. (1.5), our result
for Dy would have had the wrong sign.
(b) The x- and y-axes in Fig. 1.19b are at right angles, so it
doesn’t matter that they aren’t horizontal and vertical, respectively.
But we can’t use the angle b (the Greek letter beta) in Eqs. (1.5),
because b is measured from the +y-axis. Instead, we must use the
angle u = 90.0° - b = 90.0° - 37.0° = 53.0°. Then we find
E x = E cos 53.0° = 14.50 m21cos 53.0°2 = + 2.71 m
E y = E sin 53.0° = 14.50 m21sin 53.0°2 = + 3.59 m
evaLuaTe: Our answers to both parts are close to our predictions.
But why do the answers in part (a) correctly have only two significant figures?
using components to do vector calculations
Using components makes it relatively easy to do various calculations involving
vectors. Let’s look at three important examples: finding a vector’s magnitude and
direction, multiplying a vector by a scalar, and calculating the vector sum of two
or more vectors.
16
Chapter 1 Units, physical Quantities, and Vectors
Caution Finding the direction of a vector
from its components There’s one complication in using Eqs. (1.7) to find u: Any two
angles that differ by 180° have the same tangent. Suppose Ax = 2 m and Ay = - 2 m as
in Fig. 1.20; then tan u = - 1. But both 135°
and 315° (or - 45°) have tangents of - 1. To
decide which is correct, we have to look at
the individual components. Because Ax is
positive and Ay is negative, the angle must
be in the fourth quadrant; thus u = 315°
(or - 45°) is the correct value. Most pocket
calculators give arctan 1- 12 = - 45°. In
this case that is correct; but if instead we have
Ax = -2 m and Ay = 2 m, then the correct
angle is 135°. Similarly, when both Ax and
Ay are negative, the tangent is positive, but
the angle is in the third quadrant. Always
draw a sketch like Fig. 1.20 to determine
which of the two possibilities is correct. ❙
1.20 Drawing a sketch of a vector reveals
the signs of its x- and y-components.
Suppose that tanu =
Ay
= -1. What is u?
Ax
Two angles have tangents of -1: 135° and 315°.
The diagram shows that u must be 315°.
y
A = 2Ax2 + Ay2
Ax = 2 m
x
315°
S
A
Ay = - 2 m
1.21
Finding
the vector sum (resultant) of
S
S
A and B using components.
(1.6)
(We always take the positive root.) Equation (1.6) is valid for any choice of
x-axis and y-axis, as long as they are mutually perpendicular. The expression for the vector direction comes from the definition of the tangent of an
angle. If u is measured from the positive x-axis, and a positive angle is
measured toward the positive y-axis (as in Fig. 1.17), then
tan u =
Ay
Ax
u = arctan
and
Ay
(1.7)
Ax
We will always use the notation arctan for the inverse tangent function (see
Example 1.5 in Section 1.7). The notation tan-1 is also commonly used,
and your calculator may have an INV or 2ND button to be used with the
TAN button.
S
2. Multiplying a vector by a scalar.
If we
multiply a vector A by a scalar c,
S
S
each component of the product
D = cA is the product of c and the correS
sponding component of A:
Dx = cAx ,
135°
y
1. Finding a vector’s magnitude and direction from its components. We
can describe a vector completely by giving either its magnitude and direction or its x- and y-components. Equations (1.5) show how to find the components if we know the magnitude and direction. We can also reverse the
process: We can find the magnitude and direction if we know the components. By applying the
Pythagorean theorem to Fig. 1.17, we find that the
S
magnitude of vector A is
S
S
(components of D = cA)
Dy = cAy
(1.8)
S
For example, Eqs. (1.8) say that each component
of Sthe vector 2A is twice
S
as greatSas the corresponding component of A, so 2A is in the same direcS
tion as A but has twice the magnitude. Each component of theS vector -3A
is three times as great
as the corresponding component of
A but has the
S
S
opposite sign, so -3A is in the opposite direction from A and has three
times the magnitude. Hence Eqs. (1.8) are consistent with our discussion in
Section 1.7 of multiplying a vector by a scalar (see Fig. 1.15).
3. Using components to calculate the vectorS sum (resultant)
of two or more
S
S
vectors. Figure 1.21 shows two vectors A and B and their vector sum R,
along with the x- and y-components of all three vectors. The x-component
Rx of the vector sum is simply the sum 1Ax + Bx2 of the x-components of
the vectors being added. The same is true for the y-components. In
symbols,
S
R is the vectorSsum S
(resultant) of A and B.
S
R
By
S
Rx = Ax + Bx ,
S
S
S
Each component of R = A + B ...
Ry = Ay + By
S
B
(1.9)
S
... is the sum of the corresponding components of A and B.
Ry
Ay
O
S
A
Ax
Bx
Rx
S
The components of R Sare theSsums
of the components of A and B:
Ry = A y + By
Rx = Ax + Bx
x
Figure 1.21 shows this result for the case in which the components Ax , Ay ,
Bx , and By are all positive. Draw additional diagrams to verify
for yourself
S
S
that Eqs. (1.9) are valid for any signs of the components
of
A
and
B.
S
S
If we know the components of any two vectors A and B, perhaps
by
S
using Eqs. (1.5), we can compute the components
of
the
vector
sum
R
.
Then
S
if we need the magnitude and direction of R, we can obtain them from
Eqs. (1.6) and (1.7) with the A’s replaced by R’s.
1.8 Components of Vectors
We can
use the same procedure
to find the sum of any number ofSvecS
S S S S S
tors. If R is the vector sum of A, B, C, D, E, c, the components of R are
Rx = Ax + Bx + Cx + Dx + E x + g
Ry = Ay + By + Cy + Dy + E y + g
1.22 A vector in three dimensions.
(1.10)
In three dimensions, a vector has
x-, y-, and z-components.
z
S
A
We have talked about vectors that lie in the xy-plane only, but the component
method works just as well for vectors having any direction in space. WeS can
introduce a z-axis perpendicular to the xy-plane; then in general a vector A has
components Ax, Ay, and Az in the three coordinate directions. Its magnitude A is
A = 2Ax2 + Ay2 + Az2
17
Az
y
Ay
Ax
(1.11)
Again, we always
take the positive root (Fig. 1.22). Also, Eqs. (1.10) for
S
the vector sum R have a third component:
Rz = Az + Bz + Cz + Dz + E z + g
x
S
The magnitude of vector A
is A = 2Ax2 + Ay2 + Az2 .
We’ve focused on adding displacement vectors, but the method is applicable to all vector quantities. When we study the concept of force in Chapter 4,
we’ll find that forces are vectors that obey the same rules of vector addition.
Problem-Solving Strategy 1.3
vector addition
identify the relevant concepts: Decide what the target variable
is. It may be the magnitude of the vector sum, the direction, or
both.
the +x@axis toward the +y@axis, then its components are given
by Eqs. 1.5:
Set up the problem: Sketch the vectors being added, along with
If the angles of the vectors are given in some other way, perhaps using a different reference direction, convert them to
angles measured from the +x@axis as in Example 1.6.
2. Add the individual x-components algebraically (including
signs) to find Rx, the x-component of the vector sum. Do the
same for the y-components to find Ry. See Example 1.7.
3. Calculate the magnitude R and direction u of the vector sum by
using Eqs. (1.6) and (1.7):
suitable coordinate axes. Place the tail of the first vector at the
origin of the coordinates, place the tail of the second vectorS at the
head of the first vector, and so on. Draw the vector sum R from
the tail of the first vector (at the origin) to the head of the last
vector.
Use your sketch to estimate the magnitude and direction
S
of R. Select the mathematical tools you’ll use for the full calculation: Eqs. (1.5) to obtain the components of the vectors given,
if necessary, Eqs. (1.10) to obtain the components of the vector
sum, Eq. (1.11) to obtain its magnitude, and Eqs. (1.7) to obtain
its direction.
exeCute the solution as follows:
1. Find the x- and y-components of each individual vector and
record your results in a table, as in Example 1.7. If a vector is
described by a magnitude A and an angle u, measured from
Ax = A cos u
R = 2Rx2 + Ry2
Ry
Rx
Solution
S
u = arctan
evaluate your answer: Confirm that your results for the magnitude and direction of the vector sum agree with the estimates you
made from your sketch. The value of u that you find with a calculator may be off by 180°; your drawing will indicate the correct value.
examPle 1.7 USing comPonentS to add vectorS
Three players on a reality TV show are brought to the center of a
large, flat field. Each is given a meter stick, a compass, a calculator, a shovel, and (in a different order for each contestant) the following three displacements:
Ay = A sin u
immediately, but the winner first calculates where to go. What
does she calculate?
Solution
A: 72.4 m, 32.0° east of north
identify and Set up: The goal is to find the sum (resultant) of
S
the three displacements, so this is a problem in vector addition. See
Figure 1.23. We have chosen the +x@axis as east and the +Sy@axis as
north. We estimate from the diagram that the vector sum R is about
10 m, 40° west of north (so u is about 90° plus 40°, or about 130°).
Continued
B: 57.3 m, 36.0° south of west
S
C: 17.8 m due south
The three displacements lead to the point in the field where the
keys to a new Porsche are buried. Two players start measuring
18
Chapter 1 Units, physical Quantities, and Vectors
S
S
S
1.23 Three successive displacements
A, BS, andS C and the resulS
S
tant (vector sum) displacement R = A + B + C.
y (north)
36.0°
Distance
57.3 m
S
B
S
A
S
17.8 m C
We’ve kept an extra significant figure in the components; we’ll
round to the correct number of significant figures at the end of
our calculation. The table below shows the components of all the
displacements, the addition of the components, and the other calculations from Eqs. (1.6) and (1.7).
Angle
x-component
A = 72.4 m
58.0°
38.37 m
61.40 m
B = 57.3 m
216.0°
- 46.36 m
- 33.68 m
C = 17.8 m
270.0°
72.4 m
u = arctan
u
S
x (east)
O
0.00 m
- 17.80 m
Rx = - 7.99 m
Ry = 9.92 m
R = 21- 7.99 m22 + 19.92 m22 = 12.7 m
32.0°
R
y-component
exeCUTe: The angles of the vectors, measured from the
+ x@axis toward the + y@axis, are 190.0° - 32.0°2 = 58.0°,
1180.0° + 36.0°2 = 216.0°, and 270.0°,S respectively. We may now
use Eqs. (1.5) to find the components of A:
9.92 m
= - 51°
-7.99 m
Comparing to angle u in Fig. 1.23 shows that the calculated angle is clearly off by 180°. The correct value is u =
180° + 1- 51°2 = 129°, or 39° west of north.
evaLUaTe: Our calculated answers for R and u agree with our estimates. Notice how drawing the diagram in Fig. 1.23 made it easy to
avoid a 180° error in the direction of the vector sum.
Ax = A cos uA = 172.4 m21cos 58.0°2 = 38.37 m
Ay = A sin uA = 172.4 m21sin 58.0°2 = 61.40 m
S
S
TeST yoUr UnderSTanding
of SeCTion
1.8 Two vectors A and B lie in the
S
S
S
xy-plane. (a) Can A have the same
magnitude as B but different components? (b) Can A
S
have the same components as B but a different magnitude? ❙
1.9 UniT veCTorS
nd and ne .
1.24 (a) The unit vectors
S
(b) Expressing a vector A in terms of its
components.
(a)
Unit vectors nd and ne point in the
directions of the positive x- and y-axes
and have a magnitude of 1.
y
ne
O n
d
x
(b)
S
y
We can express a vector A in
terms of its components as
S
Ay en
A
en
O
nd
Axnd
S
A = Axnd + Ay en
x
A unit vector is a vector that has a magnitude of 1, with no units. Its only purpose is to point—that is, to describe a direction in space. Unit vectors provide a
convenient notation for many expressions involving components of vectors. We
will always include a caret, or “hat” 1^2, in the symbol for a unit vector to distinguish it from ordinary vectors whose magnitude may or may not be equal to 1.
In an xy-coordinate system we can define a unit vector nd that points in the direction of the positive x-axis and a unit vector ne that points
in the direction of the posiS
tive y-axis (Fig. 1.24a). Then we can write a vector A in terms of its components as
S
A = Ax nd + Ay ne
(1.12)
Equation (1.12) is a vector equation; each term, such as Ax nd , is a vector quantity
(Fig. 1.24b).
S
S
S
Using unit vectors, we can express the vector sum R of two vectors A and B
as follows:
S
A = Ax nd + Ay ne
S
B = Bx nd + By ne
S
S
S
R=A+B
= 1Ax nd + Ay ne 2 + 1Bx nd + By ne 2
= 1Ax + Bx2dn + 1Ay + By2 ne
= Rx nd + Ry ne
(1.13)
1.10 Products of Vectors
Equation (1.13) restates the content of Eqs. (1.9) in the form of a single vector
equation rather than two component equations.
If not all of the vectors lie in the xy-plane, then we need a third component.
We introduce a third unit vector kn that points in the direction of the positive
z-axis (Fig. 1.25). Then Eqs. (1.12) and (1.13) become
19
n.
1.25 The unit vectors nd , ne , and k
Unit vectors nd, ne, and kn point in the
directions of the positive x-, y-, and
z-axes and have a magnitude of 1.
y
en
Any vector can be expressed in terms
of its x-, y-, and z-components ...
O
S
A = Axnd + Ay ne + Az nk
(1.14)
S
B = Bx nd + By ne + Bz nk
nk
z
nd
x
n and nk.
... and unit vectors nd, e,
S
R = 1Ax + Bx2dn + 1Ay + By2en + 1Az + Bz2 kn
= R nd + R ne + R kn
x
y
(1.15)
z
Solution
ExamplE 1.8 using unit vEctors
Given the two displacements
execuTe: We have
S
S
D = 16.00 nd + 3.00 ne − 1.00kn 2 m and
F = 216.00dn + 3.00 ne − 1.00kn 2 m − 14.00dn − 5.00 ne + 8.00kn 2 m
= 3112.00 - 4.002dn + 16.00 + 5.002 ne + 1-2.00 - 8.002kn 4 m
= 18.00dn + 11.00 ne − 10.00kn 2 m
S
E = 14.00 nd − 5.00 ne + 8.00kn 2 m
S
S
find the magnitude of the displacement 2D − E.
S
From Eq. (1.11) the magnitude of F is
F = 2F x2 + F y2 + F z2
soLuTion
S
= 218.00 m22 + 111.00 m22 + 1- 10.00 m22
idenTifY and seT up: We are to multiply vector D by 2 (a scalar)
S
and
subtract
vector E from the result, so as to obtain
the vector
S
S
S
S
F = 2D − E. Equation (1.8) says that to multiply D by 2, we multiply each of its components by 2. We can use Eq. (1.15) to do the
subtraction; recall from Section 1.7 that subtracting a vector is the
same as adding the negative of that vector.
= 16.9 m
evaLuaTe: Our answer is of the same order of magnitude as the
larger components that appear in the sum. We wouldn’t expect our
answer to be much larger than this, but it could be much smaller.
TesT Your undersTanding of secTion 1.9 Arrange the following
vectors
S
in order of their magnitude,
with
the
vector
of
largest
magnitude
first.
(i)
A
=
S
S
S
(3dn + 5en − 2kn ) m; (ii) B = 1- 3dn + 5en − 2kn 2 m; (iii) C = 13dn − 5en − 2kn 2 m; (iv) D =
13dn + 5en + 2kn 2 m. ❙
1.10 producTs of vecTors
We saw how vector addition develops naturally from the problem of combining displacements. It will prove useful for calculations with many other vector
quantities. We can also express many physical relationships by using products of
vectors. Vectors are not ordinary numbers, so we can’t directly apply ordinary
multiplication to vectors. We’ll define two different kinds of products of vectors.
The first, called the scalar product, yields a result that is a scalar quantity. The
second, the vector product, yields another vector.
scalar product
We denote the scalar product of two vectors A and B by A # B. Because
ofS this
S
notation, the scalar product
is
also
called
the
dot
product.
Although
A
and
B are
S
S
vectors, the quantity A # B is a scalar.
S
S
S
S
20
ChaPter 1 Units, Physical Quantities, and Vectors
1.26 Calculating
the scalar product of two
S
S
vectors, A # B = AB cos f.
(a)
S
B
Place the vectors tail to tail.
f
S
A
To define the scalar product A # B we draw the two vectors A and B with
their tails at the same point (Fig. 1.26a). The angle f (the Greek letter phi) between their directions ranges from 0° to 180°. Figure 1.26b shows the projection
S
S
S
of vector B onto
the direction of A; this projection is the component of B in the
S
direction of A and is equal to B cos f. (We can take components
along any direcS
S
#
tion that’s Sconvenient, not just the x- and y-axes.)
We
define
A
B
to be the magS
S
nitude of A multiplied by the component of B in the direction of A, or
S
(b) A # B equals A(B cos f).
S
S
S
(Magnitude of A) * Component of B
S b
a
in direction of A
S
B
B cos f
(c) A # B also equals B(A cos f).
S
S
S
S
(Magnitude of B) * Component of A
Sb
a
in direction of B
A cos f
S
B
f
S
A
S
#
S
1.27 The scalar product A B = AB cos f
can be positive, negative, or
zero,SdependS
ing on the angle between A and B.
(a)
S
B
f
If f is between
S
S
0° and 90°, A # B
is positive ...
S
A
... because B cos f 7 0.
(b)
S
B
f
S
... because B cos f 6 0.
S
B
S
S
A
S
(1.16)
S
Alternatively, weS can define A # B to beS the magnitude of B multiplied
by
S
S
the component of A in the direction of B, as in Fig. 1.26c. Hence A # B =
B1A cos f2 = AB cos f, which is the same as Eq. (1.16).
The scalar product is a scalar quantity, not a vector, and it may be positive,
negative, or zero. When f is between 0° and 90°, cos f 7 0 and the scalar
product is positive (Fig.
1.27a). When f is Sbetween 90° and 180°
so
cos f 6 0,
S
S
S
the component of B in the directionS ofS A is negative, and A # B is negative
(Fig. 1.27b). Finally, when f = 90°, A # B = 0 (Fig. 1.27c). The scalar product
of two perpendicular vectors
is always
zero.
S
S
For
any
two
vectors
A
and
B
,
AB cos f = BA cos f. This means that
S
S
S
S
A # B = B # A. The scalar product obeys the commutative law of multiplication;
the order of the two vectors does not matter.
We’ll use the scalar product in Chapter 6 to describe work done by a force. In
later chapters we’ll use the scalar product for a variety of purposes, from calculating electric potential to determining the effects that varying magnetic fields
have on electric circuits.
S
S
using components to calculate the scalar Product
We can calculate
the scalar
product A # B directly if we know the x-, y-, and zS
S
components of A and B. To see how this is done, let’s first work out the scalar
products of the unit vectors nd , ne , and kn . All unit vectors have magnitude 1 and are
perpendicular to each other. Using Eq. (1.16), we find
S
S
nd # nd = ne # ne = kn # kn = 112112 cos 0° = 1
nd # ne = nd # kn = ne # kn = 112112 cos 90° = 0
(1.17)
S
Now we express A and B in terms of their components, expand the product, and
use these products of unit vectors:
A # B = 1Ax nd + Ay ne + Az kn 2 # 1Bx nd + By ne + Bz kn 2
S
If f = 90°,
A#B = 0
S
because B has zeroScomponent
in the direction of A.
f = 90°
S
(c)
S
Angle between A and B when placed tail to tail
S
A
Magnitudes
of
S
S
A and B
S
S
If
fSis between 90° and 180°,
S
A # B is negative ...
S
S
A ~ B = AB cos f = 0 A 0 0 B 0 cos f
S
S
A
S
Scalar (dot)
product
S
S
of vectors A and B
S
f
S
S
= Ax nd # Bx nd + Ax nd # By ne + Ax nd # Bz kn
+ Ay ne # Bx nd + Ay ne # By ne + Ay ne # Bz kn
+ Az kn # Bx nd + Az kn # By ne + Az kn # Bz kn
= Ax Bx nd # nd + Ax By nd # ne + Ax Bz nd # kn
+ Ay Bx ne # nd + Ay By ne # ne + Ay Bz ne # kn
+ Az Bx kn # nd + Az By kn # ne + Az Bz kn # kn
(1.18)
21
1.10 Products of Vectors
From Eqs. (1.17) you can see that six of these nine terms are zero. The three that
survive give
Scalar (dot)
product
S
S
of vectors A and B
S
S
Components of A
S
A ~ B = Ax Bx + Ay By + Az Bz
(1.19)
S
Components of B
Thus the scalar product of two vectors is the sum of the products of their respective components.
The scalar product
gives
a straightforward way to find the angle f between
S
S
any two vectors A and B whose components
are known. In this case we can use
S
S
Eq. (1.19) to find the scalar product of A and B. Example 1.10 shows how to do this.
Solution
ExamplE 1.9 calculating a scalar product
Find the scalar product A # B of the two vectors in Fig. 1.28. The
magnitudes of the vectors are A = 4.00 and B = 5.00.
S
S
S
f = 130.0° - 53.0° = 77.0°, so Eq. (1.16) gives us
A # B = AB cos f = 14.00215.002 cos 77.0° = 4.50
S
soLuTion
idenTifY and seT up: We can calculate the scalar product in two
ways: using the magnitudes of the vectors and the angle between
them (Eq. 1.16), and using the components of the vectors (Eq. 1.19).
We’ll do it both ways, and the results will check each other.
S
S
To use Eq. (1.19),
we must
first find the components of the vectors.
S
S
The angles of A and B are given with respect to the + x@axis and
are measured in the sense from the +x@axis to the + y@axis, so we
can use Eqs. (1.5):
Ax = 14.002 cos 53.0° = 2.407
S
1.28 Two vectors A and B in two dimensions.
Ay = 14.002 sin 53.0° = 3.195
y
Bx = 15.002 cos 130.0° = - 3.214
S
B
130.0°
By = 15.002 sin 130.0° = 3.830
As in Example 1.7, we keep an extra significant figure in the
components and round at the end. Equation (1.19) now gives us
S
A
A # B = AxBx + AyBy + AzBz
S
f
S
execuTe: The angle between the two vectors A and B is
en
53.0°
nd
x
S
= 12.40721-3.2142 + 13.195213.8302 + 102102 = 4.50
evaLuaTe: Both methods give the same result, as they should.
Solution
ExamplE 1.10 finding an anglE with thE scalar product
Find the angle between the vectors
1.29 Two vectors in three dimensions.
y
S
A = 2.00dn + 3.00 ne + 1.00kn
and
S
A extends from origin
to near corner of red box.
S
S
B = - 4.00dn + 2.00 ne − 1.00kn
B extends from origin
to far corner of blue box.
idenTifY and seT up: We’re given the x-, y-, and z-components
of two vectors. Our target variable is the angleS f Sbetween them
(Fig. 1.29). To find this, we’ll solve Eq.S(1.16),
A # B = AB cos f,
S
for f in terms of the scalar product A # B and the magnitudes
AS and
B. We can use Eq. (1.19) to evaluate the scalar product,
S
A # B = AxBx + AyBy + AzBz, and we can use Eq. (1.6) to find A
and B.
S
S
soLuTion
A
B
en
nk
z
nd
x
Continued
22
ChaPter 1 Units, Physical Quantities, and Vectors
execuTe: We solve Eq. (1.16) for cos f and use Eq. (1.19) to write
A = 2Ax2 + Ay2 + Az2 = 212.0022 + 13.0022 + 11.0022
A # B:
S
S
= 214.00
AxBx + AyBy + AzBz
A#B
=
cos f =
AB
AB
S
S
We can Suse thisS formula to find the angle between any two
vectors A and B. Here we have Ax = 2.00, Ay = 3.00, and
Az = 1.00, and Bx = - 4.00, By = 2.00, and Bz = - 1.00. Thus
B = 2Bx2 + By2 + Bz2 = 21- 4.0022 + 12.0022 + 1- 1.0022
= 221.00
cos f =
S
= 12.0021- 4.002 + 13.00212.002 + 11.0021- 1.002
= - 3.00
S
S
1.30S The
vector product of (a) A : B and
S
(b) B : A.
(a) Using theS right-hand
rule to find the
S
direction of A : B
S
S
S
1
Place A and B tail to tail.
2
Point fingers
of right hand S
S
along A, with palm facing B.
3
Curl fingers toward B.
A
4
Thumb points
in S
S
direction of A : B.
f
S
S
A:B
S
S
B
AB
- 3.00
=
214.00 221.00
f = 100°
A # B = AxBx + AyBy + AzBz
S
AxBx + AyBy + AzBz
evaLuaTe: As a check on this result, note that the scalar product
A # B is negative. This means that f is between 90° and 180° (see
Fig. 1.27), which agrees with our answer.
S
S
vector Product
S
S
We denote the
vector
product of two vectors A and B, also called the cross
S
S
product, by A : B. As the name suggests, the vector product is itself a vector.
We’ll use this product in Chapter 10 to describe torque and angular momentum;
in Chapters 27 and 28 we’ll use itS to describe
magnetic fields and forces.S
S
S
To define the vector product A : B, we again draw the two vectors A and B
with their tails at the same point (Fig. 1.30a). The two vectors then lie in a plane.
We define the vector product to be a vector quantity
with
a direction perpendicuS
S
lar to this plane (thatSis, perpendicular
to
both
A
and
B
)
and
a magnitude equal to
S
S
AB sin f. That is, if C = A : B, then
S
S
C = AB sin f
S
1
Place B and A tail to
tail.
2
Point fingersS of right
hand along B,S with
palm facing A.
3
Curl fingers toward A.
4
Thumb points in direction of B : A.
5
A
f
S
S
B
S
S
S
B:A
S
S
S
S
Magnitudes of A and B
S
S
S
S
Magnitude of vector (cross) product of vectors B and A
(b) Using theS right-hand
rule to
find the
S
S
S
direction of B : A = −A : B
(vector product is anticommutative)
S
= - 0.175
S
B : A has same magnitude as A : B
but points in opposite direction.
(1.20)
S
S
Angle between A and B
when placed tail to tail
S
We measure the angle f from A toward B and take it to be the smaller of the two
possible angles, so f ranges from 0° to 180°. Then sin f Ú 0 and C in Eq. (1.20)
S
is never
negative, as must be the case for a vector magnitude. Note that when A
S
and B are parallel or antiparallel, f = 0° or 180° and C = 0. That is, the vector
product of two parallel or antiparallel vectors is always zero. In particular, the
vector product of any vector with itself is zero.
Do not confuse the expression AB sin f for
cauTion vector product vs. scalar product
S
S
the magnitude Sof the
vector
product
A
:
B
with
the similar expression AB cos f for the
S
scalar product A # B. To see
theSdifference between these two expressions, imagineSthat we
S
S
vary the angle between A and B while keeping their magnitudes constant. When A and B
are parallel, the magnitude
of
the vector product will be zero and the scalar product will
S
S
be maximum. When A and B are perpendicular, the magnitude of the vector product will
be maximum and the scalar product will be zero. ❙
There are always two directions perpendicular to a given plane,
one on each
S
S
side of the plane. We choose
which
of
these
is
the
direction
of
A
:
B
as follows.
S
S
Imagine
rotating
vector
A
about
the
perpendicular
line
until
A
is
aligned
with
S
S
S
B, choosing the smaller of the two possible angles between A and B. Curl the
fingers of your right hand around the perpendicular line so that your fingertips
point
in the direction of rotation; your thumb will then point in the direction of
S
S
A : B. Figure 1.30a shows this right-hand rule and describes a second way to
think about this rule.
23
1.10 Products of Vectors
S
S
S
S
Similarly, we determine the direction of B : A by rotating
B into A as in
S
S
Fig. 1.30b. The result is a vector that is opposite to the vector A : B. The vector
prodS
S
uct is not commutative but instead is anticommutative: For any two vectors A and B,
S
S
S
S
A : B = −B : A
(1.21)
Just as we did for the scalar product, we can give a geometrical interpretation
of the magnitude
of the vector product. In Fig. 1.31a, B sinSf is the component of
S
vector B that isS perpendicular
to the directionSof vector A. From Eq. (1.20) the
S
magnitude
of
A
:
B
equals
the
magnitude of A multiplied by the component
of
S
S
S
S
B that is perpendicular to A.SFigure 1.31b shows that the magnitude
of
A
:
B
S
also equalsS the magnitude of B multiplied by the component of A that is perpendicular to B. Note that Fig. 1.31 shows the case in which f is between 0° and 90°;
draw a similar diagram for f between 90° and
180°
to show that the same geoS
S
metrical interpretation of the magnitude of A : B applies.
1.31 Calculating the magnitude AB sin f
of
theSvector product of two vectors,
S
A : B.
(a)
S
S
(Magnitude of A : B) equals A(B sinf).
S
S
B
S
(Magnitude of A) : Component of B S
aperpendicular to Ab
B sin f
f
S
A
(b)
S
S
(Magnitude of A : B) also equals B(A sinf).
S
using components to calculate the vector Product
S
S
If we know the components of A and B, we can calculate the components of the
vector product by using a procedure similar to that for the scalar product. First
we work out the multiplication table for unit vectors nd , ne , and kn , all three of which
are perpendicular to each other (Fig. 1.32a). The vector product of any vector
with itself is zero, so
nd : nd = ne : ne = kn : kn = 0
The boldface zero is a reminder that each product is a zero vector—that is, one
with all components equal to zero and an undefined direction. Using Eqs. (1.20)
and (1.21) and the right-hand rule, we find
nd : ne = − ne : nd = kn
ne : kn = −kn : ne = nd
S
A sinf
B
f
S
A
1.32 (a) We will always use a righthanded coordinate system, like this one.
(b) We will never use a left-handed coordinate system (in which nd : ne = −kn , and so
on).
(a) A right-handed coordinate system
y
(1.22)
kn : nd = − nd : kn = ne
en
You can verify theseSequations
by referring to Fig. 1.32a.
S
Next we express A and B in terms of their components and the corresponding
unit vectors, and we expand the expression for the vector product:
O
S
S
(Magnitude of B) : Component of A S
aperpendicular to Bb
z
nd : ne = nk
en : nk = nd
kn : nd = en
nd
nk
x
S
A : B = 1Ax nd + Ay ne + Az kn 2 : 1Bx nd + By ne + Bz kn 2
= Ax nd : Bx nd + Ax nd : By ne + Ax nd : Bz kn
(b) A left-handed coordinate system;
we will not use these.
(1.23)
+ Ay ne : Bx nd + Ay ne : By ne + Ay ne : Bz kn
en
+ Az kn : Bx nd + Az kn : By ne + Az kn : Bz kn
We can also rewrite the individual terms in Eq. (1.23) as Ax nd : By ne =
1Ax By2 nd : ne , and so on. Evaluating these by using the multiplication table for the
unit vectors in Eqs. (1.22) and then grouping the terms, we get
S
S
A : B = 1Ay Bz - Az By2 nd + 1Az Bx - Ax Bz2 ne + 1Ax By - Ay Bx2kn
(1.24)
If
youS compare
Eq. (1.24) with Eq. (1.14), you’ll see that the components of
S
S
C = A : B are
S
S
Components of vector (cross) product A : B
Cx = Ay Bz - Az By
Cy = Az Bx - Ax Bz
S
Ax , Ay , Az = components of A
y
Cz = Ax By - Ay Bx
S
Bx , By , Bz = components of B
(1.25)
z
nk
O
nd
x
24
ChaPter 1 Units, Physical Quantities, and Vectors
With the axis system of Fig. 1.32a, if we reverse the direction of the z-axis, we
get the system shown in Fig. 1.32b. Then, as you may verify, the definition of the
vector product gives nd : ne = −kn instead of nd : ne = kn . In fact, all vector products of unit vectors nd , ne , and kn would have signs opposite to those in Eqs. (1.22).
So there are two kinds of coordinate systems, which differ in the signs of the vector products of unit vectors. An axis system in which nd : ne = kn , as in Fig. 1.32a,
is called a right-handed system. The usual practice is to use only right-handed
systems, and we’ll follow that practice throughout this book.
Solution
ExamplE 1.11 calculating a vEctor product
S
Vector A has magnitude
6 units and is in the direction of the
S
+ x@axis. Vector B has magnitude 4 units and lies in the xy-plane,
making an angle
of 30°
with the + x@axis (Fig. 1.33). Find the vecS
S
S
tor product C = A : B.
soLuTion
idenTifY and seT up: We’ll find the vector product in two ways,
which will provide a check of our calculations. First we’ll use
Eq. (1.20) and the right-hand rule; then we’ll use Eqs. (1.25) to find
the vector product by using components.
S
S
S
S
S
1.33 Vectors
A and B and their vector product C = A : B.
S
Vector B lies in the xy-plane.
y
f = 30°
A
S
z
C
S
S
By the right-hand rule, the direction of AS: B Sis along
the
S
+z@axis (the direction of the unit vector kn ), so C = A : B = 12knS.
ToS use Eqs. (1.25),
we first determine the components of A
S
and B. Note that A points
along the x-axis, so its only nonzero
S
component is Ax. For B, Fig. 1.33 shows that f = 30° is measured
from the +x-axis toward the +y-axis, so we can use Eqs. (1.5):
Ax = 6
Ay = 0
Az = 0
Bx = 4 cos 30° = 223
By = 4 sin 30° = 2
Bz = 0
Then Eqs. (1.25) yield
Cy = 10212232 - 162102 = 0
S
S
AB sin f = 1621421sin 30°2 = 12
Cx = 102102 - 102122 = 0
B
O
execuTe: From Eq. (1.20) the magnitude of the vector product is
x
Cz = 162122 - 10212232 = 12
S
Thus again we have C = 12kn .
evaLuaTe: Both methods give the same result. Depending on the
situation, one or the other of the two approaches may be the more
convenient one to use.
S
TesT Your
undersTanding of secTion 1.10
Vector
A has magnitude 2
S
S
S
and vector B has magnitude 3. The angle f between A and B is (i) 0°, (ii) 90°, or
(iii) 180°. For each of the following situations, state what the value Sof f
must be.
S
#
(In each
situation
there
may
be
more
than
one
correct
answer.)
(a)
A
B
= 0;
S
S
S
S
S
S
S
S
(b) A : B = 0; (c) A # B = 6; (d) A # B = - 6; (e) 1magnitude of A : B2 = 6. ❙
ChaPter
1 sUmmary
SolutionS to all exampleS
Physical quantities and units: Three fundamental
physical quantities are mass, length, and time. The
corresponding fundamental SI units are the kilogram, the meter, and the second. Derived units for
other physical quantities are products or quotients
of the basic units. Equations must be dimensionally
consistent; two terms can be added only when they
have the same units. (See Examples 1.1 and 1.2.)
Significant figures: The accuracy of a measure-
Significant figures in magenta
ment can be indicated by the number of significant
figures or by a stated uncertainty. The significant
figures in the result of a calculation are determined
by the rules summarized in Table 1.2. When only
crude estimates are available for input data, we can
often make useful order-of-magnitude estimates.
(See Examples 1.3 and 1.4.)
C
0.424 m
=
= 3.14
2r
210.06750 m2
p =
123.62 + 8.9 = 132.5
Scalars, vectors, and vector addition: Scalar quantities are numbers and combine according to the
usual rules of arithmetic. Vector quantities have direction as well as magnitude and combine according to the rules of vector addition. The negative of
a vector has the same magnitude but points in the
opposite direction. (See Example 1.5.)
S
S
A
Vector components and vector addition: Vectors
Rx = Ax + Bx
can be added by usingS components
of vectors.
S
S
The x-component Sof R =S A + B is the sum of the
x-components of A and B, and likewise for the
y- and z-components. (See Examples 1.6 and 1.7.)
Ry = Ay + By
+
S
=
B
S
A+ B
S
A
S
B
y
S
(1.9)
Rz = Az + Bz
Ry
R
By
S
B
Ay
S
A
O
Ax
x
Bx
Rx
Unit vectors: Unit vectors describe directions in
space. A unit vector has a magnitude of 1, with no
units. The unit vectors nd , ne , and kn , aligned with the
x-, y-, and z-axes of a rectangular coordinate system, are especially useful. (See Example 1.8.)
Scalar product:
TheSscalar product C = A # B of
S
S
S
two vectors A and B is a scalar quantity. It can
S
be
expressed in terms of the magnitudes of A and
S
B and the angle f between the
twoSvectors, or in
S
terms of the components Sof A
and
B.SThe scalar
S
S
product is commutative; A # B = B # A. The scalar
product of two perpendicular vectors is zero. (See
Examples 1.9 and 1.10.)
S
A = Ax nd + Ay ne + Az kn
(1.14)
y
Ay en
S
A
en
O
A # B = AB cos f = 0 A 0 0 B 0 cos f
S
S
S
S
A # B = Ax Bx + Ay By + Az Bz
S
S
(1.16)
(1.19)
nd
= Axnd + Ay ne
x
Axnd
Scalar product A # B = AB cosf
S
S
S
B
f
S
A
25
26
ChaPter 1 Units, Physical Quantities, and Vectors
S
S
S
Vector product: SThe vector
product C = SA : B
S
of two vectors SA and
B is a third vector C. The
S
magnitude
of
A
:
B
depends
on the magnitudes
S
S
of A and B and the angle
f
between
the two vecS
S
tors. The direction of A : B is perpendicular to
the plane of the two vectors being multiplied, as
given
bySthe right-hand
rule. The components
S
S
of C = A : B can
be
expressed
in terms of the
S
S
components of A and
BS. The vector
product is
S
S
S
not commutative; A : B = −B : A. The vector
product of two parallel or antiparallel vectors is
zero. (See Example 1.11.)
(1.20)
(1.25)
Cz = Ax By - Ay Bx
S
A
f
S
B
S
S
(Magnitude of A : B) = AB sinf
vEctors on thE roof
An air-conditioning unit is fastened to a roof that slopes at an
angle
S
of 35° above the horizontal (Fig. 1.34). Its weight is a force F on the
air conditioner that is directed vertically downward. In order that
the unit not crush the roof tiles, the component of the unit’s weight
perpendicular to the roof cannot exceed 425 N. (One newton, or
1 N, is the SI unit of force. It is equal to 0.2248 lb.) (a) What is the
maximum allowed weight of the unit? (b) If the fasteners fail, the
unit slides 1.50 m along the roof before it comes to a halt against a
ledge. How much work does the weight force do on the unit during
its slide if the unit
has the weight calculated in part (a)? The work
S
S
done by a force F on an object that undergoes a displacement s is
S S
W = F # s.
S
A : B is perpendicular
S
S
to the plane of A and B.
S
A:B
Cx = Ay Bz - Az By
Cy = Az Bx - Ax Bz
S
S
Solution
Bridging proBlEm
C = AB sin f
1.34 An air-conditioning unit on a slanted roof.
1.50 m
y
S
F
35°
x
soLuTion guide
idenTifY and seT up
1. This problem involves vectors and components. What are the
known quantities? Which aspect(s) of the weight vector (magnitude, direction, and/or particular components) represent the
target variable for part (a)? Which aspect(s) must you know to
solve part (b)?
2. Make a sketch based on Fig. 1.34. Draw the x- and y-axes,
choosing the positive direction for each. Your axes don’t have
to be horizontal and vertical, but they do have to be mutually
perpendicular. Figure 1.34 shows a convenient choice of axes:
The x-axis is parallel to the slope of the roof.
3. Choose the equations you’ll use to determine the target
variables.
execuTe
4. Use the relationship between the magnitude and direction of a
vector and its components to solve for the target variable in
part (a). Be careful: Is 35° the correct angle to use in the equation? (Hint: Check your sketch.)
5. Make sure your answer has the correct number of significant
figures.
6. Use the definition of the scalar product to solve for the target
variable in part (b). Again, use the correct number of significant figures.
evaLuaTe
7. Did your answer to part (a) include a vector component whose
absolute value is greater than the magnitude of the vector? Is
that possible?
8. There are two ways to find the scalar product of two vectors,
one of which you used to solve part (b). Check your answer by
repeating the calculation, using the other way. Do you get the
same answer?
exercises
27
Problems
For assigned homework and other learning materials, go to MasteringPhysics®.
., .., ...: Difficulty levels. CP: Cumulative problems incorporating material from earlier chapters. CALC: Problems requiring calculus.
DATA: Problems involving real data, scientific evidence, experimental design, and/or statistical reasoning. bio: Biosciences problems.
diSCUSSion QUeSTionS
Q1.1 How many correct experiments do we need to disprove a
theory? How many do we need to prove a theory? Explain.
Q1.2 Suppose you are asked to compute the tangent of
5.00 meters. Is this possible? Why or why not?
Q1.3 What is your height in centimeters? What is your weight in
newtons?
Q1.4 The U.S. National Institute of Standards and Technology
(NIST) maintains several accurate copies of the international
standard kilogram. Even after careful cleaning, these national
standard kilograms are gaining mass at an average rate of about
1 mg>y (y = year) when compared every 10 years or so to the
standard international kilogram. Does this apparent increase have
any importance? Explain.
Q1.5 What physical phenomena (other than a pendulum or cesium clock) could you use to define a time standard?
Q1.6 Describe how you could measure the thickness of a sheet of
paper with an ordinary ruler.
Q1.7 The quantity p = 3.14159 c is a number with no dimensions, since it is a ratio of two lengths. Describe two or three other
geometrical or physical quantities that are dimensionless.
Q1.8 What are the units of volume? Suppose another student
tells you that a cylinder of radius r and height h has volume given
by pr 3h. Explain why this cannot be right.
Q1.9 Three archers each fire four arrows at a target. Joe’s four
arrows hit at points 10 cm above, 10 cm below, 10 cm to the left,
and 10 cm to the right of the center of the target. All four of Moe’s
arrows hit within 1 cm of a point 20 cm from the center, and Flo’s
four arrows hit within 1 cm of the center. The contest judge says
that one of the archers is precise but not accurate, another archer
is accurate but not precise, and the third archer is both accurate
and precise. Which description applies to which archer? Explain.
Q1.10 Is the vector 1dn + ne + kn 2 a unit vector? Is the vector
13.0dn − 2.0en2 a unit vector? Justify your answers.
Q1.11 A circular racetrack has a radius of 500 m. What is the
displacement of a bicyclist when she travels around the track from
the north side to the south side? When she makes one complete
circle around the track? Explain.
Q1.12 Can you find two vectors with different lengths that have
a vector sum of zero? What length restrictions are required for
three vectors to have a vector sum of zero? Explain.
Q1.13 The “direction of time” is said to proceed from past to
future. Does this mean that time is a vector quantity? Explain.
Q1.14 Air traffic controllers give instructions called “vectors”
to tell airline pilots in which direction they are to fly. If these are
the only instructions given, is the name “vector” used correctly?
Why or why not?
Q1.15 Can you find a vector quantity that has a magnitude
of zero but components that are not zero? Explain. Can the
magnitude of a vector be less than the magnitude of any of its
components? Explain.
Q1.16 (a) Does it make sense to say that a vector is negative?
Why? (b) Does it make sense to say that one vector is the negative
of another? Why? Does your answer here contradict what you said
in part (a)?
S
S
S
Q1.17 If C = SA + BS, what must be true about the directions and
magnitudes of A and B if C = AS+ B? SWhat must be true about
the directions
and magnitudes
of A and B if C = 0?
S
S
Q1.18
If SA and
B are nonzero vectors, is it possible for both
S
S
S
A ~ B and A : B to be
zero?
Explain.
S
S
Q1.19 What does A ~SA, the
scalar product of a vector with itS
self, give? What about A : A, the vector product of a vector with
itself?
S
S
Q1.20 Let A represent any nonzero vector. Why is A >SA a unit
vector, and what is its direction? SIf u is the angle that A makes
with the +x-axis, explain why 1A >A2 ~ nd is called the direction
cosine for that axis.
Q1.21 Figure 1.7 shows the result of an unacceptable error in the
stopping position of a train. If a train travels 890 km from Berlin
to Paris and then overshoots the end of the track by 10.0 m, what is
the percent error in the total distance covered? Is it correct to write
the total distance covered by the train as 890,010 m? Explain.
Q1.22 Which Sof the
following
are Slegitimate
mathematical
S
S
S
S
S
S
S
operations:
(a)
A
~
1B
−
C
2;
(b)
1A
−
B
2
:
C
;
(c)
A ~ 1B : C2;
S
S
S
S
S
S
(d) A : 1B : C2; (e) A : 1B ~ C2? In each case, give the reason
for your answer.
S
S
S
Q1.23
Consider
the vector products A : 1B : C2 and
S
S
S
1A : B2 : C. Give an example that illustrates the general rule
that these two vector products do not
have theS same magnitude or
S S
direction. Can you choose vectors A, B, and C such that these two
vector products are equal? If so, give
an example.
S
S
S
S
S
Q1.24 Show that, no matter what A and B are, A ~ 1A : B2 = 0.
(Hint: Do not look for an elaborate mathematical proof. Consider
the definition of
the
direction of the cross product.)
S
S
Q1.25 (a) If A ~ B = 0,
does
it necessarily follow that A = 0 or
S
S
B = 0? Explain. (b) If A : B = 0, does it necessarily follow that
A = 0 or BS = 0 ? Explain.
Q1.26 If A = 0 for a vector in the xy-plane, does it follow that
Ax = - Ay ? What can you say about Ax and Ay?
exerCiSeS
Section 1.3 Standards and Units
Section 1.4 Using and Converting Units
1.1 . Starting with the definition 1 in. = 2.54 cm, find the number of (a) kilometers in 1.00 mile and (b) feet in 1.00 km.
1.2 .. According to the label on a bottle of salad dressing, the
volume of the contents is 0.473 liter (L). Using only the conversions 1 L = 1000 cm3 and 1 in. = 2.54 cm, express this volume
in cubic inches.
1.3 .. How many nanoseconds does it take light to travel 1.00 ft
in vacuum? (This result is a useful quantity to remember.)
1.4 .. The density of gold is 19.3 g>cm3. What is this value in
kilograms per cubic meter?
1.5 . The most powerful engine available for the classic 1963
Chevrolet Corvette Sting Ray developed 360 horsepower and had
a displacement of 327 cubic inches. Express this displacement in
liters (L) by using only the conversions 1 L = 1000 cm3 and
1 in. = 2.54 cm.
28
Chapter 1 Units, physical Quantities, and Vectors
Section 1.6 estimates and Orders of Magnitude
1.17
..
bio A rather ordinary middle-aged man is in the hospital for a routine checkup. The nurse writes “200” on the patient’s
..
S
y
For the vectors A and Figure e1.24
S
B (15.0 m)
B in Fig. E1.24, use a scale
drawing to find the magnitude
and direction
of (a) the vector
S
S
S
30.0°
D (10.0 m)
sum A + B Sand S(b) the vector
difference A − B. Use your
53.0°
answers to find the magnitude
S
S
x
and direction
ofS (c) −A − B
O
S
25.0°
and (d) B − A. (See also
Exercise 1.31 for a different
S
S
C (12.0 m)
approach.)
A (8.00 m)
..
1.25
A postal employee
drives a delivery truck along the
route shown in Fig. E1.25. Determine the magnitude and direction of the resultant displacement by drawing a scale diagram.
(See also Exercise 1.32 for a different approach.)
1.24
S
Figure e1.25
km
STOP
1
. With a wooden ruler, you measure the length of a rectangular piece of sheet metal to be 12 mm. With micrometer calipers,
you measure the width of the rectangle to be 5.98 mm. Use the
correct number of significant figures: What is (a) the area of
the rectangle; (b) the ratio of the rectangle’s width to its length;
(c) the perimeter of the rectangle; (d) the difference between the
length and the width; and (e) the ratio of the length to the width?
1.15 .. A useful and easy-to-remember approximate value for the
number of seconds in a year is p * 107. Determine the percent
error in this approximate value. (There are 365.24 days in one year.)
1.16 . Express each approximation of p to six significant
figures: (a) 22>7 and (b) 355>113. (c) Are these approximations
accurate to that precision?
1.14
Section 1.7 Vectors and Vector addition
3.
Section 1.5 Uncertainty and Significant Figures
medical chart but forgets to include the units. Which of these
quantities could the 200 plausibly represent? The patient’s
(a) mass in kilograms; (b) height in meters; (c) height in centimeters; (d) height in millimeters; (e) age in months.
1.18 . How many gallons of gasoline are used in the United
States in one day? Assume that there are two cars for every three
people, that each car is driven an average of 10,000 miles per year,
and that the average car gets 20 miles per gallon.
1.19 . bio How many times does a typical person blink her
eyes in a lifetime?
1.20 . bio Four astronauts are in a spherical space station. (a) If,
as is typical, each of them breathes about 500 cm3 of air with each
breath, approximately what volume of air (in cubic meters) do these
astronauts breathe in a year? (b) What would the diameter (in meters) of the space station have to be to contain all this air?
1.21 . In Wagner’s opera Das Rheingold, the goddess Freia is
ransomed for a pile of gold just tall enough and wide enough to
hide her from sight. Estimate the monetary value of this pile. The
density of gold is 19.3 g>cm3, and take its value to be about
$10 per gram.
1.22 . bio How many times does a human heart beat during a
person’s lifetime? How many gallons of blood does it pump?
(Estimate that the heart pumps 50 cm3 of blood with each beat.)
1.23 . You are using water to dilute small amounts of chemicals in
the laboratory, drop by drop. How many drops of water are in a 1.0-L
bottle? (Hint: Start by estimating the diameter of a drop of water.)
4.0 km
2.6 km
1.6 .. A square field measuring 100.0 m by 100.0 m has an area
of 1.00 hectare. An acre has an area of 43,600 ft2. If a lot has an
area of 12.0 acres, what is its area in hectares?
1.7 . How many years older will you be 1.00 gigasecond from
now? (Assume a 365-day year.)
1.8 . While driving in an exotic foreign land, you see a speed limit
sign that reads 180,000 furlongs per fortnight. How many miles per
hour is this? (One furlong is 18 mile, and a fortnight is 14 days. A furlong originally referred to the length of a plowed furrow.)
1.9 . A certain fuel-efficient hybrid car gets gasoline mileage of
55.0 mpg (miles per gallon). (a) If you are driving this car in
Europe and want to compare its mileage with that of other
European cars, express this mileage in km>L 1L = liter2. Use the
conversion factors in Appendix E. (b) If this car’s gas tank holds
45 L, how many tanks of gas will you use to drive 1500 km?
1.10 . The following conversions occur frequently in physics
and are very useful. (a) Use 1 mi = 5280 ft and 1 h = 3600 s to
convert 60 mph to units of ft>s. (b) The acceleration of a freely
falling object is 32 ft>s2. Use 1 ft = 30.48 cm to express this acceleration in units of m>s2. (c) The density of water is 1.0 g>cm3.
Convert this density to units of kg>m3.
1.11 .. Neptunium. In the fall of 2002, scientists at Los
Alamos National Laboratory determined that the critical mass of
neptunium-237 is about 60 kg. The critical mass of a fissionable
material is the minimum amount that must be brought together to
start a nuclear chain reaction. Neptunium-237 has a density of
19.5 g>cm3. What would be the radius of a sphere of this material
that has a critical mass?
1.12 . bio (a) The recommended daily allowance (RDA) of the
trace metal magnesium is 410 mg>day for males. Express this
quantity in mg>day. (b) For adults, the RDA of the amino acid
lysine is 12 mg per kg of body weight. How many grams per day
should a 75-kg adult receive? (c) A typical multivitamin tablet can
contain 2.0 mg of vitamin B2 (riboflavin), and the RDA is
0.0030 g>day. How many such tablets should a person take each
day to get the proper amount of this vitamin, if he gets none from
other sources? (d) The RDA for the trace element selenium is
0.000070 g>day. Express this dose in mg>day.
1.13 .. bio Bacteria. Bacteria vary in size, but a diameter of
2.0 mm is not unusual. What are the volume (in cubic centimeters)
and surface area (in square millimeters) of a spherical bacterium
of that size? (Consult Appendix B for relevant formulas.)
START
45°
N
W
E
S
.. A spelunker is surveying a cave. She follows a passage
180 m straight west, then 210 m in a direction 45° east of south,
and then 280 m at 30° east of north. After a fourth displacement,
1.26
Problems
she finds herself back where she started. Use a scale drawing to
determine the magnitude and direction of the fourth displacement. (See also Problem 1.61 for a different approach.)
Section 1.8 Components of Vectors
1.27S . Compute the x- and y-components of the vectors A, B, C,
S
S
S
and D in Fig. E1.24.
S
1.28 .. Let u be the angle that the vector A makes with the
+ x-axis, measured counterclockwise from that axis. Find
angle u for a vector that has these components: (a) Ax = 2.00 m,
Ay = - 1.00 m; (b) Ax = 2.00 m, Ay = 1.00 m; (c) Ax = - 2.00 m,
Ay = 1.00 m; (d)S Ax = - 2.00 m, Ay = - 1.00 m.
S
1.29 . Vector A has y-component Ay = + 9.60 m. A makes an
angle of 32.0° counterclockwise
from the +y-axis.S (a) What is the
S
x-component of A
?
(b)
What
is
the
magnitude of A?
S
1.30 . Vector A is in the direction
34.0° clockwise from the
S
- y@axis. The x-component
of
A
is
A
=
- 16.0 m.S (a) What is the
x
S
y-component of A? (b) What
is
the
magnitude
of A?
S
S
1.31 . For the vectors A and B in Fig. E1.24, use the method of
components
to find the magnitude
andS direction of (a) the vector
S
S
S
sum
A
+
B
;
(b)
the
vector
sum
B
+
A
; (c) the vector difference
S
S
S
S
A − B; (d) the vector difference B − A.
1.32 .. A postal employee drives a delivery truck over the route
shown in Fig. E1.25. Use the method of components to determine
the magnitude and direction of her resultant displacement. In a
vector-addition diagram (roughly to scale), show that the resultant
displacement found from your diagram is in qualitative agreement
with the result you obtained by using the method of components.
1.33 .. A disoriented physics professor drives 3.25 km north,
then 2.20 km west, and then 1.50 km south. Find the magnitude
and direction of the resultant displacement, using the method of
components. In a vector-addition diagram (roughly to scale), show
that the resultant displacement found from your diagram is in
qualitative agreement with the result you obtained by using the
method of components.
1.34 . Find the magnitude and direction of the vector represented by the following pairs of components: (a) Ax = -8.60 cm,
Ay = 5.20 cm; (b) Ax = - 9.70 m, Ay = -2.45 m; (c) Ax =
7.75 km, Ay = - 2.70
km.
S
1.35 .. Vector A is 2.80 cm Figure E1.35
long and is 60.0° above the
y
x-axis Sin the first quadrant.
S
Vector B is 1.90 cm long and is
A (2.80 cm)
60.0° below the x-axis in the
fourth quadrant (Fig. E1.35).
Use components to find the
mag60.0°
S
S
x
nitude
and
direction
of
(a)
A
+
B
;
S
S
S
S
O
60.0°
(b) A − B; (c) B − A. In each
S
case, sketch the vector addition or
B (1.90 cm)
subtraction and show that your
numerical answers are in qualitative agreement with your sketch.
Section 1.9 Unit Vectors
1.36S
.
S
In each case, findSthe x- and y-components
of vector A:
S
(a) A = 5.0dn − 6.3en; (b) A = 11.2en − 9.91dn; (c) A = -15.0dn +
S
S
S
22.4en ; (d) A = 5.0B, where B = 4dn − 6en.
1.37 .. Write each vector in Fig. E1.24 in terms of the unit vectors nd and ne .
S
S
1.38 .. Given two vectors A = 4.00dn + 7.00en and B = 5.00dn −
2.00en, (a) find the magnitude of each vector; (b) use unit vectors
S
29
S
to write an expression for the vector difference A − BS; andS(c) find
the magnitude and direction
of the Svector
difference A − B. (d) In
S S
S
a vector diagram show A, B, and A − B, and show that your diagram agrees qualitatively with your answer to part (c).
1.39 .. (a) Write each vector in Figure E1.39
Fig. E1.39 in terms of the unit
y
vectors nd and ne . (b) Use Sunit vecS
A (3.60 m)
tors
to express
vector
C, where
S
S
S
C = 3.00 A − 4.00 B. (c) Find
theS magnitude and direction
70.0°
of C.
x
1.40 S. You are given two vecO
30.0°
tors
A = -3.00dn + 6.00en and
S
S
B (2.4 m)
B = 7.00dn + 2.00en. Let counterclockwise angles be
positive.
S
(a)
What angle does A make with the +x@axis?
(b) What angle
does
S
S
S
S
B make
with
the
+x@axis?
(c)
Vector
C
is
the
sum
of
A
and
B,
S
S
S
S
so C = A + B. What angle does
C make with the + x@axis?
. Given two vectors AS = - 2.00dn + 3.00en + 4.00kn and
1.41
S
B = 3.00dn + 1.00en − 3.00kn , (a) find the magnitude of each vector; (b) use
unit
vectors to write an expression for the vector difS
S
ference
A
−
B
;
and
(c) find the magnitude of the
vector
difference
S
S
S
S
A − B. Is this the same as the magnitude of B − A? Explain.
Section 1.10 Products of Vectors
1.42
..
S
S
(a) Find the scalar product of the vectors A and B given
in Exercise 1.38. (b) FindStheS angle Sbetween these two vectors.
1.43 . For the
vectors
A,SB, andS C Sin Fig. E1.24, find the scalar
S
S
S
products (a) A ~ B; (b) B ~ C; (c) A ~S C. S
1.44 .. Find the vector product A : B (expressed in unit vectors) of the two vectors given in Exercise 1.38. What is the magnitude of the vector product?
1.45 .. Find the angle between each of these pairs of vectors:
S
S
(a) A = - 2.00dn + 6.00en
and
B = 2.00dn − 3.00en
(b) A = 3.00dn + 5.00en
and
B = 10.00dn + 6.00en
(c) A = - 4.00dn + 2.00en
and
B = 7.00dn + 14.00en
S
S
S
S
. For the two vectors in Fig. E1.35, find the magnitude and
S
S
direction
of (a) the vector product A : B; (b) the vector product
S
S
B : A.
S
S
1.47 . For the two vectors A and D in Fig. E1.24,
find the magS
S
nitude and direction
of
(a)
the
vector
product
A
:
D
; (b) the vecS
S
tor product D : A.
S
S
1.48 . For theS two
vectors A and B in Fig. E1.39, find (a) the
S
scalar product
A
~
B
;
(b) the magnitude and direction of the vector
S
S
product A : B.
1.46
probLems
.. White Dwarfs and Neutron Stars. Recall that density
is mass divided by volume, and consult Appendix B as needed.
(a) Calculate the average density of the earth in g>cm3, assuming
our planet is a perfect sphere. (b) In about 5 billion years, at
the end of its lifetime, our sun will end up as a white dwarf that
has about the same mass as it does now but is reduced to about
15,000 km in diameter. What will be its density at that stage?
(c) A neutron star is the remnant of certain supernovae (explosions
of giant stars). Typically, neutron stars are about 20 km in diameter and have about the same mass as our sun. What is a typical
neutron star density in g>cm3?
1.49
30
Chapter 1 Units, physical Quantities, and Vectors
.
An acre has a length of one furlong 118 mi2 and a width
one-tenth of its length. (a) How many acres are in a square mile?
(b) How many square feet are in an acre? See Appendix E. (c) An
acre-foot is the volume of water that would cover 1 acre of flat
land to a depth of 1 foot. How many gallons are in 1 acre-foot?
1.51 .. An Earthlike Planet. In January 2006 astronomers reported the discovery of a planet, comparable in size to the earth,
orbiting another star and having a mass about 5.5 times the earth’s
mass. It is believed to consist of a mixture of rock and ice, similar
to Neptune. If this planet has the same density as Neptune
11.76 g>cm32, what is its radius expressed (a) in kilometers and
(b) as a multiple of earth’s radius? Consult Appendix F for astronomical data.
1.52 .. The Hydrogen Maser. A maser is a laser-type device
that produces electromagnetic waves with frequencies in the microwave and radio-wave bands of the electromagnetic spectrum.
You can use the radio waves generated by a hydrogen maser as a
standard of frequency. The frequency of these waves is
1,420,405,751.786 hertz. (A hertz is another name for one cycle
per second.) A clock controlled by a hydrogen maser is off by only
1 s in 100,000 years. For the following questions, use only three
significant figures. (The large number of significant figures given
for the frequency simply illustrates the remarkable accuracy to
which it has been measured.) (a) What is the time for one cycle of
the radio wave? (b) How many cycles occur in 1 h? (c) How many
cycles would have occurred during the age of the earth, which is
estimated to be 4.6 * 109 years? (d) By how many seconds would
a hydrogen maser clock be off after a time interval equal to the
age of the earth?
1.53 . bio Breathing Oxygen. The density of air under standard laboratory conditions is 1.29 kg>m3, and about 20% of that
air consists of oxygen. Typically, people breathe about 12 L of air
per breath. (a) How many grams of oxygen does a person breathe
in a day? (b) If this air is stored uncompressed in a cubical tank,
how long is each side of the tank?
1.54 ... A rectangular piece of aluminum is 7.60 { 0.01 cm
long and 1.90 { 0.01 cm wide. (a) Find the area of the rectangle
and the uncertainty in the area. (b) Verify that the fractional uncertainty in the area is equal to the sum of the fractional uncertainties in the length and in the width. (This is a general result.)
1.55 ... As you eat your way through a bag of chocolate chip
cookies, you observe that each cookie is a circular disk with a diameter of 8.50 { 0.02 cm and a thickness of 0.050 { 0.005 cm.
(a) Find the average volume of a cookie and the uncertainty in the
volume. (b) Find the ratio of the diameter to the thickness and the
uncertainty in this ratio.
1.56 . bio Biological tissues are typically made up of 98% water.
Given that the density of water is 1.0 * 103 kg>m3, estimate the
mass of (a) the heart of an adult human; (b) a cell with a diameter
of 0.5 mm; (c) a honeybee.
1.57 . bio Estimate the number of atoms in your body. (Hint:
Based on what you know about biology and chemistry, what are
the most common types of atom in your body? What is the mass
of each type of atom? Appendix D gives the atomic masses of different elements, measured in atomic mass units; you can find the
value of an atomic mass unit, or 1 u, in Appendix E.)
1.58 .. Two ropes in a vertical plane exert equal-magnitude
forces on a hanging weight but pull with an angle of 72.0° between them. What pull does each rope exert if their resultant pull
is 372 N directly upward?
1.59 ... Two workers pull horizontally on a heavy box, but one
pulls twice as hard as the other. The larger pull is directed at 21.0°
1.50
west of north, and the resultant of these two pulls is 460.0 N directly northward. Use vector components to find the magnitude of
each of these pulls and the direction of the smaller pull.
1.60 .. Three horizontal ropes Figure P1.60
pull on a large stone stuck in the
y
S
ground, S producing
the vector
S
S
B (80.0 N)
forces A, B, and C shown in
S
A (100.0 N)
30.0°
Fig. P1.60. Find the magnitude
and direction of a fourth force
on the stone that will make the
30.0°
vector sum of the four forces
x
O
53.0°
zero.
1.61 ... As noted in Exercise
S
C (40.0 N)
1.26, a spelunker is surveying a
cave. She follows a passage 180 m
straight west, then 210 m in a direction 45° east of south, and then
280 m at 30° east of north. After a fourth displacement, she finds
herself back where she started. Use the method of components to
determine the magnitude and direction of the fourth displacement. Draw the vector-addition diagram and show that it is in
qualitative agreement with your numerical solution.
1.62 ... Emergency Landing. A plane leaves the airport in
Galisteo and flies 170 km at 68.0° east of north; then it changes
direction to fly 230 km at 36.0° south of east, after which it makes
an immediate emergency landing in a pasture. When the airport
sends out a rescue crew, in which direction and how far should
this crew fly to go directly to this plane?
1.63 ... bio Dislocated Shoulder. A patient with a dislocated
shoulder isSput into
a traction apparatus as shown in Fig. P1.63.
S
The pulls A and B have equal magnitudes and must combine to
produce an outward traction force of 12.8 N on the patient’s arm.
How large should these pulls be?
Figure P1.63
S
A
32°
32°
S
B
1.64
..
A sailor in a small sailboat encounters shifting winds.
She sails 2.00 km east, next 3.50 km southeast, and then an additional distance in an unknown direction. Her final position is
5.80 km directly east of the starting point (Fig. P1.64). Find the
magnitude and direction of the third leg of the journey. Draw the
vector-addition diagram and show that it is in qualitative agreement with your numerical solution.
Figure P1.64
N
W
E
START
FINISH
5.80 km
S
2.00 km
3.50 km
45.0°
Third
leg
problems
.. You leave the airport in College Station and fly 23.0 km
in a direction 34.0o south of east. You then fly 46.0 km due north.
How far and in what direction must you then fly to reach a private
landing strip that is 32.0 km due west of the College Station
airport?
1.66 ... On a training flight, a Figure P1.66
student pilot flies from Lincoln,
IOWA
Nebraska, to Clarinda, Iowa,
147 km
next to St. Joseph, Missouri, and
Clarinda
Lincoln 85°
then to Manhattan, Kansas
(Fig. P1.66). The directions are
106 km
NEBRASKA
shown relative to north: 0° is
167°
north, 90° is east, 180° is south,
St. Joseph
and 270° is west. Use the method
Manhattan
of components to find (a) the dis166 km
tance she has to fly from
235°
N
Manhattan to get back to
W
E
Lincoln, and (b) the direction
S KANSAS
MISSOURI
(relative to north) she must fly to
get there. Illustrate your solutions with a vector diagram.
1.67 .. As a test of orienteering skills, your physics class holds a
contest in a large, open field. Each contestant is told to travel 20.8 m
due north from the starting point, then 38.0 m due east, and finally 18.0 m in the direction 33.0° west of south. After the specified displacements, a contestant will find a silver dollar hidden
under a rock. The winner is the person who takes the shortest time
to reach the location of the silver dollar. Remembering what you
learned in class, you run on a straight line from the starting point
to the hidden coin. How far and in what direction do you run?
1.68 ... Getting Back. An explorer in Antarctica leaves his
shelter during a whiteout. He takes 40 steps northeast, next
80 steps at 60° north of west, and then 50 steps due south. Assume
all of his steps are equal in length. (a) Sketch, roughly to scale, the
three vectors and their resultant. (b) Save the explorer from becoming hopelessly lost by giving him the displacement, calculated
by using the method of components, that will return him to his
shelter.
1.69 .. You are lost at night in a large, open field. Your GPS tells
you that you are 122.0 m from your truck, in a direction 58.0° east
of south. You walk 72.0 m due west along a ditch. How much farther, and in what direction, must you walk to reach your truck?
1.70 ... A ship leaves the island of Guam and sails 285 km at
62.0° north of west. In which direction must it now head and how
far must it sail so that its resultant displacement will be 115 km
directly east of Guam?
1.71 .. bio Bones and Muscles. A physical therapy patient has
a forearm that weighs 20.5 N and lifts a 112.0-N weight. These
two forces are directed vertically downward. The only other significant forces on this forearm come from the biceps muscle
(which acts perpendicular to the forearm) and the force at the
elbow. If the biceps produces a pull of 232 N when the forearm is
raised 43.0° above the horizontal, find the magnitude and direction of the force that the elbow exerts on the forearm. (The sum of
the elbow force and the biceps force must balance the weight of
the arm and the weight it is carrying, so their vector sum must be
132.5 N, upward.)
1.72 ... You decide to go to your favorite neighborhood restaurant. You leave your apartment, take the elevator 10 flights down
(each flight is 3.0 m), and then walk 15 m south to the apartment
exit. You then proceed 0.200 km east, turn north, and walk
0.100 km to the entrance of the restaurant. (a) Determine the displacement from your apartment to the restaurant. Use unit vector
1.65
31
notation for your answer, clearly indicating your choice of coordinates. (b) How far did you travel along the path you took from
your apartment to the restaurant, and what is the magnitude of the
displacement you calculated in part (a)?
1.73 .. While following a treasure map, you start at an old oak
tree. You first walk 825 m directly south, then turn and walk
1.25 km at 30.0° west of north, and finally walk 1.00 km at 32.0°
north of east, where you find the treasure: a biography of Isaac
Newton! (a) To return to the old oak tree, in what direction should
you head and how far will you walk? Use components to solve this
problem. (b) To see whether your calculation in part (a) is reasonable, compare it with a graphical solution drawn roughly to scale.
1.74 .. A fence post is 52.0 m from where you are standing, in a
direction 37.0° north of east. A second fence post is due south
from you. How far are you from the second post if the distance
between the two posts is 68.0 m?
1.75 .. A dog in an open field runs 12.0 m east and then 28.0 m
in a direction 50.0° west of north. In what direction and how far
must the dog then run to end up 10.0 m south of her original starting point?
1.76 ... Ricardo and Jane are standing under a tree in the middle of a pasture. An argument ensues, and they walk away in different directions. Ricardo walks 26.0 m in a direction 60.0° west
of north. Jane walks 16.0 m in a direction 30.0° south of west.
They then stop and turn to face each other. (a) What is the distance between them? (b) In what direction should Ricardo walk to
go directly toward Jane?
1.77 ... You are camping with Joe and Karl. Since all three of
you like your privacy, you don’t pitch your tents close together.
Joe’s tent is 21.0 m from yours, in the direction 23.0° south of east.
Karl’s tent is 32.0 m from yours, in the direction 37.0° north of
east. What is the distance between Karl’s tent and Joe’s tent?
1.78 .. Bond Angle in Methane. In the methane molecule,
CH4, each hydrogen atom is at a corner of a regular tetrahedron
with the carbon atom at the center. In coordinates for which one of
the C ¬ H bonds is in the direction of nd + ne + kn , an adjacent
C ¬ H bond is in the nd − ne − kn direction. Calculate the angle between these two bonds.
S
S
1.79 .. Vectors A and B have scalar product - 6.00, and their
vector product has magnitude +9.00. What is the angle between
these two vectors?
1.80 .. A cube is placed so Figure P1.80
that one corner is at the origin
z
and three edges are along the
x-, y-, and z-axes of a coordinate
b
c
system (Fig. P1.80). Use vectors
to compute (a) the angle bed
tween the edge along the z-axis
(line ab) and the diagonal from
a
y
the origin to the opposite corner
(line ad), and (b) the angle between line ac (the diagonal of a
x
face) and line ad.S
S
1.81 .. Vector A has magnitude 12.0
m, and vector B has magS
S
nitude 16.0 m. The scalar product A ~ B is 112.0 m2. What is the
magnitude of the vector product between these two vectors?
1.82 ... Obtain a unit vector perpendicular to the two vectors
given in Exercise 1.41.
S
S
1.83 ..S The scalar product of vectors A and B is + 48.0 m2.
Vector A has
magnitude 9.00 m and direction 28.0° west of south.
S
If vectorS B has direction 39.0° south of east, what is the magnitude of B?
32
Chapter 1 Units, physical Quantities, and Vectors
..
S
S
Two vectors A and B have magnitudes
A = 3.00 and
S
S
n + 2.00dn.
B = 3.00. Their vector product
is
A
:
B
=
-5.00k
S
S
What is the angle between A and B? S
S
1.85 .. You are given vectors
A = 5.0dn − 6.5en and BS =
S
3.5dn − 7.0en. A third vector,
C, lies in the xy-plane. Vector
C
is
S
S
S
perpendicular to vector A, and the scalar product of C with
B
is
S
15.0. From this information, find the components of vector C.
1.86 .. Later in our study of physics we will encounter quantiS
S
S
ties represented by 1A : B2 ~ C. (a) Prove that for any three vecS
S
S
S
S
S
S
S
S
tors
AS, B,S and C, A ~ 1B
: C2 = 1A : B2 ~ C. (b) Calculate
S
S
1A : B2 ~ C for vector A with magnitude A = 5.00 and angle
uA =S 26.0° (measured from the +x-axis toward the S+y-axis), vector B with B = 4.00 and uB = 63.0°, and vector
C Swith magniS
tude 6.00 and in the + z-direction. Vectors A and B are in the
xy-plane.
1.87 ... DATA You are a team leader at a pharmaceutical company. Several technicians are preparing samples, and you want to
compare the densities of the samples (density = mass>volume) by
using the mass and volume values they have reported. Unfortunately,
you did not specify what units to use. The technicians used a variety of units in reporting their values, as shown in the following
table.
1.84
Sample ID
Mass
Volume
A
8.00 g
1.67 * 10-6 m3
B
6.00 mg
9.38 * 106 mm3
C
8.00 mg
2.50 * 10-3 cm3
D
9.00 * 10-4 kg
2.81 * 103 mm3
E
9.00 * 104 ng
1.41 * 10-2 mm3
F
6.00 * 10-2 mg
1.25 * 108 mm3
List the sample IDs in order of increasing density of the sample.
1.88 ... DATA You are a mechanical
engineer
working for a
S
S
manufacturing company. Two forces, F1 and F2, act on a component part of a piece of equipment. Your boss asked you to find the
magnitude of the
larger
of these two forces. You can vary the
S
S
angle between F1 and F2 from 0o to 90o while the magnitude of
each force stays constant. And, you can measure the magnitude of
the resultant force they produce (their vector sum), but you cannot
directly measure the magnitude of each separate force. You measure the magnitude of the resultant force for four angles u between
the directions of the two forces as follows:
U
Resultant force (N)
0.0°
8.00
45.0°
7.43
60.0°
7.00
90.0°
5.83
(a) What is the magnitude of the larger of the two forces?
(b) When the equipment is used on the production line, the angle
between the two forces is 30.0°. What is the magnitude of the
resultant force in this case?
1.89 ... DATA Navigating in the Solar System. The Mars
Polar Lander spacecraft was launched on January 3, 1999. On
December 3, 1999, the day Mars Polar Lander impacted the
Martian surface at high velocity and probably disintegrated, the
positions of the earth and Mars were given by these coordinates:
x
y
z
Earth
0.3182 AU
0.9329 AU
0.0000 AU
Mars
1.3087 AU
- 0.4423 AU
- 0.0414 AU
With these coordinates, the sun is at the origin and the earth’s
orbit is in the xy-plane. The earth passes through the + x-axis once
a year on the autumnal equinox, the first day of autumn in the
northern hemisphere (on or about September 22). One AU, or
astronomical unit, is equal to 1.496 * 108 km, the average distance from the earth to the sun. (a) Draw the positions of the sun,
the earth, and Mars on December 3, 1999. (b) Find these distances
in AU on December 3, 1999: from (i) the sun to the earth; (ii) the
sun to Mars; (iii) the earth to Mars. (c) As seen from the earth,
what was the angle between the direction to the sun and the direction to Mars on December 3, 1999? (d) Explain whether Mars was
visible from your current location at midnight on December 3,
1999. (When it is midnight, the sun is on the opposite side of the
earth from you.)
ChaLLenge ProBLeMS
1.90
...
Completed Pass. The football team at Enormous State
University (ESU) uses vector displacements to record its plays,
with the origin taken to be the position of the ball before the play
starts. In a certain pass play, the receiver starts at + 1.0dn − 5.0en,
where the units are yards, nd is to the right, and ne is downfield.
Subsequent displacements of the receiver are + 9.0dn (he is in motion before the snap), +11.0en (breaks downfield), - 6.0dn + 4.0en
(zigs), and +12.0dn + 18.0en (zags). Meanwhile, the quarterback
has dropped straight back to a position -7.0en. How far and in
which direction must the quarterback throw the ball? (Like the
coach, you will be well advised to diagram the situation before
solving this numerically.)
1.91 ... Navigating in the Big Dipper. All of the stars of the
Big Dipper (part of the constellation Ursa Major) may appear to be
the same distance from the earth, but in fact they are very far from
each other. Figure P1.91 shows the distances from the earth to
each of these stars. The distances are given in light-years (ly), the
distance that light travels in one year. One light-year equals
9.461 * 1015 m. (a) Alkaid and Merak are 25.6o apart in the
earth’s sky. In a diagram, show the relative positions of Alkaid,
Merak, and our sun. Find the distance in light-years from Alkaid
to Merak. (b) To an inhabitant of a planet orbiting Merak, how
many degrees apart in the sky would Alkaid and our sun be?
Figure P1.91
Megrez
81 ly
Mizar
73 ly
Alkaid
138 ly
Alioth
64 ly
Dubhe
105 ly
Merak
77 ly
Phad
80 ly
answers
BIO CalCulaTing lung VolumE in Humans. In humans, oxygen and carbon dioxide are exchanged in the blood
within many small sacs called alveoli in the lungs. Alveoli provide a large surface area for gas exchange. Recent careful measurements show that the total number of alveoli in a typical pair of
lungs is about 480 * 106 and that the average volume of a single
alveolus is 4.2 * 106 mm3. (The volume of a sphere is V = 43 pr 3,
and the area of a sphere is A = 4pr 2.)
1.92 What is total volume of the gas-exchanging region of the
lungs? (a) 2000 mm3; (b) 2 m3; (c) 2.0 L; (d) 120 L.
1.93 If we assume that alveoli are spherical, what is the diameter
of a typical alveolus? (a) 0.20 mm; (b) 2 mm; (c) 20 mm;
(d) 200 mm.
1.94 Individuals vary considerably in total lung volume.
Figure P1.94 shows the results of measuring the total lung volume and average alveolar volume of six individuals. From these
data, what can you infer about the relationship among alveolar
size, total lung volume, and number of alveoli per individual? As
the total volume of the lungs increases, (a) the number and volume
of individual alveoli increase; (b) the number of alveoli increases
and the volume of individual alveoli decreases; (c) the volume of
the individual alveoli remains constant and the number of alveoli
increases; (d) both the number of alveoli and the volume of
individual alveoli remain constant.
Figure P1.94
Average alveolar volume
passage probLems
33
6
5
4
3
2
1
0
500 1000 1500 2000 2500
Total lung volume
Answers
chapter opening Question
?
(iii) Take the + x@axis to point east and the +y@axis to point north.
Then we need to find the y-component of the velocity vector,
which has magnitude v = 15 km>h and is at an angle u = 37°
measured from the + x@axis toward the +y@axis. From Eqs. (1.5)
we have vy = v sin u = 115 km>h2 sin 37° = 9.0 km>h. So the
thunderstorm moves 9.0 km north in 1 h and 18 km north in 2 h.
Test Your understanding Questions
-4
3
3
3
1.5 (ii) Density = 11.80 kg2>16.0 * 10 m 2 = 3.0 * 10 kg>m .
When we multiply or divide, the number with the fewest significant figures controls the number of significant figures in the
result.
1.6 The answer depends on how many students are enrolled at
your campus.
S
1.7 (ii),S (iii), Sand S(iv) S Vector S−T has the same magnitude as
vector T, so S − T = S + 1−T2 is the sum of one vector of
magnitude 3 m
and one
of magnitude 4 m. This sum has SmagS
S
nitude
7 m if S and −T are parallel and
magnitude
1m
if S and
S
S
S
S
S
−T are antiparallel. The magnitude
of
S
−
T
is
5
m
if
S
and
−T
S S
S
S
are perpendicular, when vectors S, T, and S − T form a 3–4–5
right triangle. Answer (i) is impossible because the magnitude
of the sum of two vectors cannot be greater than the sum of the
magnitudes; answer (v) is impossible because the sum of two vectors can be zero only if the two vectors are antiparallel and have
the same magnitude; and answer (vi) is impossible because the
magnitude of a vector cannot be negative.
S
S
1.8 (a) yes, (b) no Vectors A and B can have the same magnitude but different components if they point in different directions. If they Shave Sthe same components, however, they are the
same vector 1A = B2 and so must have the sameSmagnitude.
S
S
S
1.9 All have the same magnitude. Vectors A, B, C, and D
point in different directions but have the same magnitude:
A = B = C = D = 21{3 m22 + 1{5 m22 + 1{2 m22
= 29 m2 + 25 m2 + 4 m2 = 238 m2 = 6.2 m
1.10 (a) (ii) F = 90°, (b) (i) F = 0° or (iii) F = 180°,
(c) (i) F = 0°, (d) (iii) F =S 180°, S(e) (ii) F = 90° (a) The scalar product is zero only if A and
B Sare perpendicular. (b) The
S
vector product is zero only if A and B are parallel or antiparallel. (c) The
scalar product isSequal Sto the product of the magniS
S
tudes 1A # B = AB2 only if A and B are parallel. (d) The scalar
productSis Sequal to the negative
of the
product of the magniS
S
tudes 1A # B = -AB2 only if A and B are antiparallel. (e) The
magnitude of the vector product
is equal to the product
of
the
S
S
S
S
magnitudes 31magnitude of A : B2 = AB4 only if A and B are
perpendicular.
bridging problem
(a) 5.2 * 102 N
(b) 4.5 * 102 N # m
?
A typical runner gains
speed gradually during
the course of a sprinting foot
race and then slows down
after crossing the finish line. In
which part of the motion is it
accurate to say that the runner
is accelerating? (i) During the
race; (ii) after the runner crosses
the finish line; (iii) both (i) and (ii);
(iv) neither (i) nor (ii); (v) answer
depends on how rapidly the
runner gains speed during
the race.
2
Motion Along A
StrAight line
Learning goaLs
Looking forward at …
2.1 How the ideas of displacement and average
2.2
2.3
2.4
2.5
2.6
velocity help us describe straight-line
motion.
The meaning of instantaneous velocity;
the difference between velocity and speed.
How to use average acceleration and instantaneous acceleration to describe changes in
velocity.
How to use equations and graphs to solve
problems that involve straight-line motion
with constant acceleration.
How to solve problems in which an object is
falling freely under the influence of gravity
alone.
How to analyze straight-line motion when
the acceleration is not constant.
Looking back at …
1.7 The displacement vector.
1.8 Components of a vector.
W
hat distance must an airliner travel down a runway before it reaches
takeoff speed? When you throw a baseball straight up in the air, how
high does it go? When a glass slips from your hand, how much time
do you have to catch it before it hits the floor? These are the kinds of questions
you will learn to answer in this chapter. We begin our study of physics with
mechanics, the study of the relationships among force, matter, and motion. In this
chapter and the next we will study kinematics, the part of mechanics that enables
us to describe motion. Later we will study dynamics, which helps us understand
why objects move in different ways.
In this chapter we’ll concentrate on the simplest kind of motion: a body moving along a straight line. To describe this motion, we introduce the physical quantities velocity and acceleration. In physics these quantities have definitions that
are more precise and slightly different from the ones used in everyday language.
Both velocity and acceleration are vectors: As you learned in Chapter 1, this
means that they have both magnitude and direction. Our concern in this chapter
is with motion along a straight line only, so we won’t need the full mathematics
of vectors just yet. But using vectors will be essential in Chapter 3 when we consider motion in two or three dimensions.
We’ll develop simple equations to describe straight-line motion in the important special case when acceleration is constant. An example is the motion of a
freely falling body. We’ll also consider situations in which acceleration varies
during the motion; in this case, it’s necessary to use integration to describe the
motion. (If you haven’t studied integration yet, Section 2.6 is optional.)
2.1 DispLacement, time, anD average veLocity
Suppose a drag racer drives her dragster along a straight track (Fig. 2.1). To
study the dragster’s motion, we need a coordinate system. We choose the x-axis to
lie along the dragster’s straight-line path, with the origin O at the starting line.
34
2.1 Displacement, Time, and Average Velocity
2.1 Positions of a dragster at two times during its run.
START
O
Position at t1 = 1.0 s
Position at t2 = 4.0 s
P1
P2
x-axis
x1 = 19 m
Displacement from t1 to t2
∆x = 1x2 - x12 = 258 m
x-coordinate of
dragster at 1.0 s
x is positive to the right of the
origin (O), negative to the left
of it.
FINISH
x
x2 = 277 m
x-coordinate of
dragster at 4.0 s
When the dragster moves in the +x-direction, the displacement
∆x is positive and so is the average x-velocity:
∆x 258 m
vav-x =
=
= 86 m>s
∆t
3.0 s
We also choose a point on the dragster, such as its front end, and represent the
entire dragster by that point. Hence we treat the dragster as a particle.
A useful way to describe the motion of this particle is in terms of the change
in its coordinate x over a time interval. Suppose that 1.0 s after the start the front
of the dragster is at point P1, 19 m from the origin, and 4.0 s after the start it is
at point P2, 277 m from the origin. The displacement of the particle is a vector
that points from P1 to P2 (see Section 1.7). Figure 2.1 shows that this vector points
along the x-axis. The x-component (see Section 1.8) of the displacement is the
change in the value of x, (277 m - 19 m) = 258 m, that took place during the
time interval of (4.0 s - 1.0 s) = 3.0 s. We define the dragster’s average velocity
during this time interval as a vector whose x-component is the change in x divided by the time interval: (258 m)>(3.0 s) = 86 m>s.
In general, the average velocity depends on the particular time interval chosen. For a 3.0-s time interval before the start of the race, the dragster is at rest at
the starting line and has zero displacement, so its average velocity for this time
interval is zero.
Let’s generalize the concept of average velocity. At time t1 the dragster is at
point P1, with coordinate x1, and at time t2 it is at point P2, with coordinate x2.
The displacement of the dragster during the time interval from t1 to t2 is the
vector from P1 to P2. The x-component of the displacement, denoted ∆x, is the
change in the coordinate x:
∆x = x2 - x1
(2.1)
The dragster moves along the x-axis only, so the y- and z-components of the displacement are equal to zero.
caution The meaning of 𝚫x Note that ∆x is not the product of ∆ and x; it is a single
symbol that means “the change in quantity x.” We use the Greek capital letter ∆ (delta) to
represent a change in a quantity, equal to the final value of the quantity minus the initial
value—never the reverse. Likewise, the time interval from t1 to t2 is ∆t, the change in t:
∆t = t2 - t1 (final time minus initial time). ❙
The x-component of average velocity, or the average x-velocity, is the
x-component of displacement, ∆x, divided by the time interval ∆t during which
the displacement occurs. We use the symbol vav@x for average x-velocity (the subscript “av” signifies average value, and the subscript x indicates that this is the
x-component):
Average x-velocity of a
particle in straight-line
motion during time
interval from t1 to t2
x-component of the particle’s displacement
vav-x =
Time interval
x - x1
∆x
= 2
t2 - t1
∆t
Final x-coordinate
minus initial
x-coordinate
Final time minus initial time
(2.2)
35
36
Chapter 2 Motion along a Straight Line
2.2 Positions of an official’s truck at two
times during its motion. The points P1 and
P2 now indicate the positions of the truck,
not the dragster, and so are the reverse of
Fig. 2.1.
START
Position at t2 = 25.0 s
Position at t1 = 16.0 s
P1
P2
FINISH
Displacement from t1 to t2
O
x2 = 19 m
This position is now x2.
x
x1 = 277 m
∆x = 1x2 - x12 = - 258 m
This position is now x1.
When the truck moves in the - x-direction, ∆x is
negative and so is the average x-velocity:
∆x
- 258 m
vav-x =
=
= - 29 m>s
9.0 s
∆t
As an example, for the dragster in Fig. 2.1, x1 = 19 m, x2 = 277 m, t1 = 1.0 s,
and t2 = 4.0 s. So Eq. (2.2) gives
vav@x =
Rules for the Sign
Table 2.1 of x-Velocity
If x-coordinate is:
. . . x-velocity is:
Positive & increasing
(getting more positive)
Positive: Particle
is moving in
+ x-direction
Positive & decreasing
(getting less positive)
Negative: Particle
is moving in
- x-direction
Negative & increasing
(getting less negative)
Positive: Particle
is moving in
+ x-direction
Negative & decreasing
(getting more negative)
Negative: Particle
is moving in
- x-direction
Note: These rules apply to both the average
x-velocity vav@x and the instantaneous x-velocity
vx (to be discussed in Section 2.2).
277 m - 19 m
258 m
=
= 86 m>s
4.0 s - 1.0 s
3.0 s
The average x-velocity of the dragster is positive. This means that during the
time interval, the coordinate x increased and the dragster moved in the positive
x-direction (to the right in Fig. 2.1).
If a particle moves in the negative x-direction during a time interval, its average velocity for that time interval is negative. For example, suppose an official’s
truck moves to the left along the track (Fig. 2.2). The truck is at x1 = 277 m at
t1 = 16.0 s and is at x2 = 19 m at t2 = 25.0 s. Then ∆x = 119 m - 277 m2 =
-258 m and ∆t = (25.0 s - 16.0 s) = 9.0 s. The x-component of average velocity is vav@x = ∆x>∆t = (-258 m)>(9.0 s) = -29 m>s. Table 2.1 lists some
simple rules for deciding whether the x-velocity is positive or negative.
Caution The sign of average x-velocity In our example positive vav@x means motion to
the right, as in Fig. 2.1, and negative vav@x means motion to the left, as in Fig. 2.2. But that’s
only because we chose the +x-direction to be to the right. Had we chosen the + x-direction
to be to the left, the average x-velocity vav@x would have been negative for the dragster
moving to the right and positive for the truck moving to the left. In most problems the
direction of the coordinate axis is yours to choose. Once you’ve made your choice, you
must take it into account when interpreting the signs of vav@x and other quantities that
describe motion! ❙
With straight-line motion we sometimes call ∆x simply the displacement and
vav@x simply the average velocity. But remember that these are the x-components of
vector quantities that, in this special case, have only x-components. In Chapter 3,
displacement, velocity, and acceleration vectors will have two or three nonzero
components.
Figure 2.3 is a graph of the dragster’s position as a function of time—that
is, an x-t graph. The curve in the figure does not represent the dragster’s path;
2.3 A graph of the position of a dragster
as a function of time.
Dragster track
(not to scale)
P2
x (m) For a displacement along the x-axis, an object’s average x-velocity
vav-x equals the slope of a line connecting the corresponding points
400
on a graph of position (x)
versus time (t).
300
x2
elo
-v
200
x
ge
100
P1
x1
O
p2
y
cit
o
pe
Sl
p1
1
t1
=
era
av
∆x = x2 - x1
Slope = rise over run =
∆t = t2 - t1
2
3
4
t2
5
t (s)
∆x
∆t
2.2 Instantaneous Velocity
as Fig. 2.1 shows, the path is a straight line. Rather, the graph represents how
the dragster’s position changes with time. The points p1 and p2 on the graph
correspond to the points P1 and P2 along the dragster’s path. Line p1 p2 is the
hypotenuse of a right triangle with vertical side ∆x = x2 - x1 and horizontal
side ∆t = t2 - t1 . The average x-velocity vav@x = ∆x >∆t of the dragster equals
the slope of the line p1 p2 —that is, the ratio of the triangle’s vertical side ∆x to its
horizontal side ∆t. (The slope has units of meters divided by seconds, or m>s, the
correct units for average x-velocity.)
The average x-velocity depends on only the total displacement ∆x = x2 - x1
that occurs during the time interval ∆t = t2 - t1 , not on what happens during
the time interval. At time t1 a motorcycle might have raced past the dragster at
point P1 in Fig. 2.1, then slowed down to pass through point P2 at the same time t2
as the dragster. Both vehicles have the same displacement during the same time
interval and so have the same average x-velocity.
If distance is given in meters and time in seconds, average velocity is measured in meters per second, or m>s (Table 2.2). Other common units of velocity
are kilometers per hour (km>h), feet per second (ft>s), miles per hour (mi>h), and
knots (1 knot = 1 nautical mile>h = 6080 ft>h).
37
typical velocity
Table 2.2 magnitudes
A snail’s pace
A brisk walk
Fastest human
Freeway speeds
Fastest car
Random motion of air
molecules
Fastest airplane
10-3 m>s
2 m>s
11 m>s
30 m>s
341 m>s
500 m>s
1000 m>s
Orbiting communications
satellite
Electron orbiting in a
hydrogen atom
2 * 106 m>s
Light traveling in vacuum
3 * 108 m>s
3000 m>s
test your unDerstanDing of section 2.1 Each of the following five trips
takes one hour. The positive x-direction is to the east. (i) Automobile A travels 50 km due
east. (ii) Automobile B travels 50 km due west. (iii) Automobile C travels 60 km due east,
then turns around and travels 10 km due west. (iv) Automobile D travels 70 km due
east. (v) Automobile E travels 20 km due west, then turns around and travels 20 km
due east. (a) Rank the five trips in order of average x-velocity from most positive to most
negative. (b) Which trips, if any, have the same average x-velocity? (c) For which trip, if
any, is the average x-velocity equal to zero? ❙
2.2 instantaneous veLocity
Sometimes average velocity is all you need to know about a particle’s motion.
For example, a race along a straight line is really a competition to see whose average velocity, vav@x , has the greatest magnitude. The prize goes to the competitor
who can travel the displacement ∆x from the start to the finish line in the shortest time interval, ∆t (Fig. 2.4).
But the average velocity of a particle during a time interval can’t tell us how
fast, or in what direction, the particle was moving at any given time during the
interval. For that we need to know the instantaneous velocity, or the velocity at
a specific instant of time or specific point along the path.
caution How long is an instant? You might use the phrase “It lasted just an instant”
to refer to something that spanned a very short time interval. But in physics an instant has
no duration at all; it refers to a single value of time. ❙
To find the instantaneous velocity of the dragster in Fig. 2.1 at point P1 , we
move point P2 closer and closer to point P1 and compute the average velocity
vav@x = ∆x>∆t over the ever-shorter displacement and time interval. Both ∆x
and ∆t become very small, but their ratio does not necessarily become small.
In the language of calculus, the limit of ∆x>∆t as ∆t approaches zero is called
the derivative of x with respect to t and is written dx>dt. We use the symbol vx ,
with no “av” subscript, for the instantaneous velocity along the x-axis, or the
instantaneous x-velocity:
The instantaneous
x-velocity of a particle in
straight-line motion ...
dx
∆x
=
∆t S 0 ∆t
dt
vx = lim
... equals the limit of the particle’s average
x-velocity as the time interval approaches zero ...
(2.3)
... and equals the instantaneous rate of
change of the particle’s x-coordinate.
2.4 The winner of a 50-m swimming race
is the swimmer whose average velocity
has the greatest magnitude—that is, the
swimmer who traverses a displacement ∆x
of 50 m in the shortest elapsed time ∆t.
38
Chapter 2 Motion along a Straight Line
2.5 In any problem involving straight-line
motion, the choice of which direction is
positive and which is negative is entirely
up to you.
A bicyclist moving to the left ...
O
... has a negative x-velocity vx if we choose
the positive x-direction to the right ...
x
O
... but has a positive x-velocity vx if we
choose the positive x-direction to the left.
x
The time interval ∆t is always positive, so vx has the same algebraic sign as
∆x. A positive value of vx means that x is increasing and the motion is in the
positive x-direction; a negative value of vx means that x is decreasing and the
motion is in the negative x-direction. A body can have positive x and negative
vx , or the reverse; x tells us where the body is, while vx tells us how it’s moving
(Fig. 2.5). The rules that we presented in Table 2.1 (Section 2.1) for the sign of
average x-velocity vav@x also apply to the sign of instantaneous x-velocity vx .
Instantaneous velocity, like average velocity, is a vector; Eq. (2.3) defines its
x-component. In straight-line motion, all other components of instantaneous
velocity are zero. In this case we often call vx simply the instantaneous velocity.
(In Chapter 3 we’ll deal with the general case in which the instantaneous velocity
can have nonzero x-, y-, and z-components.) When we use the term “velocity,” we
will always mean instantaneous rather than average velocity.
“Velocity” and “speed” are used interchangeably in everyday language, but
they have distinct definitions in physics. We use the term speed to denote distance traveled divided by time, on either an average or an instantaneous basis.
Instantaneous speed, for which we use the symbol v with no subscripts, measures how fast a particle is moving; instantaneous velocity measures how fast
and in what direction it’s moving. Instantaneous speed is the magnitude of instantaneous velocity and so can never be negative. For example, a particle with
instantaneous velocity vx = 25 m>s and a second particle with vx = -25 m>s
are moving in opposite directions at the same instantaneous speed 25 m>s.
Caution Average speed and average velocity Average speed is not the magnitude of
average velocity. When César Cielo set a world record in 2009 by swimming 100.0 m in
46.91 s, his average speed was (100.0 m)>(46.91 s) = 2.132 m>s. But because he swam
two lengths in a 50-m pool, he started and ended at the same point and so had zero total
displacement and zero average velocity! Both average speed and instantaneous speed are
scalars, not vectors, because these quantities contain no information about direction. ❙
Solution
ExAmplE 2.1 AvErAgE And instAntAnEous vElocitiEs
A cheetah is crouched 20 m to the east of a vehicle (Fig. 2.6a). At
time t = 0 the cheetah begins to run due east toward an antelope
that is 50 m to the east of the vehicle. During the first 2.0 s
of the chase, the cheetah’s x-coordinate varies with time according
to the equation x = 20 m + (5.0 m>s2)t 2. (a) Find the cheetah’s
displacement between t1 = 1.0 s and t2 = 2.0 s. (b) Find its average velocity during that interval. (c) Find its instantaneous velocity at t1 = 1.0 s by taking ∆t = 0.1 s, then 0.01 s, then 0.001 s.
(d) Derive an expression for the cheetah’s instantaneous velocity as
a function of time, and use it to find vx at t = 1.0 s and t = 2.0 s.
2.6 A cheetah attacking an antelope from ambush. The animals are not drawn to the same scale as the axis.
(a) The situation
(b) Our sketch
(c) Decisions
1 Point axis in
direction cheetah runs,
so that all values will
be positive.
2 Place origin
at vehicle.
3 Mark initial
positions of cheetah
and antelope.
4 Mark positions
for cheetah at 1 s
and 2 s.
5 Add the known
and unknown
quantities.
2.2 Instantaneous Velocity
The average x-velocity during this 0.1-s interval is
soLution
iDentiFy and set up: Figure 2.6b shows our sketch of the cheetah’s motion. We use Eq. (2.1) for displacement, Eq. (2.2) for average
velocity, and Eq. (2.3) for instantaneous velocity.
vav@x =
x1 and x2 are
x1 = 20 m + 15.0 m>s2211.0 s22 = 25 m
x2 = 20 m + 15.0 m>s2212.0 s22 = 40 m
The displacement during this 1.0-s interval is
∆x = x2 - x1 = 40 m - 25 m = 15 m
vx =
(b) The average x-velocity during this interval is
vav@x
26.05 m - 25 m
= 10.5 m>s
1.1 s - 1.0 s
Following this pattern, you can calculate the average x-velocities
for 0.01-s and 0.001-s intervals: The results are 10.05 m>s and
10.005 m>s. As ∆t gets smaller, the average x-velocity gets closer
to 10.0 m>s, so we conclude that the instantaneous x-velocity at
t = 1.0 s is 10.0 m>s. (We suspended the rules for significant-figure
counting in these calculations.)
(d) From Eq. (2.3) the instantaneous x-velocity is vx = dx>dt.
The derivative of a constant is zero and the derivative of t 2 is 2t, so
execute: (a) At t1 = 1.0 s and t2 = 2.0 s the cheetah’s positions
dx
d
=
320 m + 15.0 m>s22t 24
dt
dt
= 0 + 15.0 m>s2212t2 = 110 m>s22t
x2 - x1
40 m - 25 m
15 m
=
=
=
t2 - t1
2.0 s - 1.0 s
1.0 s
At t = 1.0 s, this yields vx = 10 m>s, as we found in part (c); at
t = 2.0 s, vx = 20 m>s.
= 15 m>s
evaLuate: Our results show that the cheetah picked up speed from
t = 0 (when it was at rest) to t = 1.0 s 1vx = 10 m>s2 to t = 2.0 s
1vx = 20 m>s2. This makes sense; the cheetah covered only 5 m
during the interval t = 0 to t = 1.0 s, but it covered 15 m during the
interval t = 1.0 s to t = 2.0 s.
(c) With ∆t = 0.1 s the time interval is from t1 = 1.0 s to a
new t2 = 1.1 s. At t2 the position is
x2 = 20 m + 15.0 m>s2211.1 s22 = 26.05 m
Finding velocity on an x-t graph
We can also find the x-velocity of a particle from the graph of its position as a
function of time. Suppose we want to find the x-velocity of the dragster in Fig. 2.1
at point P1 . As point P2 in Fig. 2.1 approaches point P1 , point p2 in the x-t graphs
of Figs. 2.7a and 2.7b approaches point p1 and the average x-velocity is calculated
over shorter time intervals ∆t. In the limit that ∆t S 0, shown in Fig. 2.7c, the
slope of the line p1 p2 equals the slope of the line tangent to the curve at point p1 .
Thus, on a graph of position as a function of time for straight-line motion, the
instantaneous x-velocity at any point is equal to the slope of the tangent to the
curve at that point.
If the tangent to the x-t curve slopes upward to the right, as in Fig. 2.7c, then
its slope is positive, the x-velocity is positive, and the motion is in the positive
x-direction. If the tangent slopes downward to the right, the slope of the x-t graph
and the x-velocity are negative, and the motion is in the negative x-direction.
When the tangent is horizontal, the slope and the x-velocity are zero. Figure 2.8
(next page) illustrates these three possibilities.
2.7 Using an x-t graph to go from (a), (b) average x-velocity to (c) instantaneous x-velocity vx. In (c) we find the
slope of the tangent to the x-t curve by dividing any vertical interval (with distance units) along the tangent by the
corresponding horizontal interval (with time units).
(a)
x (m)
400
(b)
x (m)
400
∆t = 2.0 s
∆x = 150 m
vav-x = 75 m>s
p2
300
200
100
O
39
300
200
∆t
1
2
3
x (m)
400
∆t = 1.0 s
∆x = 55 m
vav-x = 55 m>s
p1
4
5
t (s)
As the average x-velocity vav-x is calculated
over shorter and shorter time intervals ...
O
1
vx =
200
inst
100
p1
∆t ∆x
2
160 m
4.0 s
= 40 m>s
300
p2
100
∆x
p1
(c)
3
4
5
t (s)
... its value vav-x = ∆x>∆t approaches the
instantaneous x-velocity.
O
1
=
ent
ang locity
t
f
pe o -ve
Slo neous x
a
t
n
a
160 m
4.0 s
2
3
4
5
t (s)
The instantaneous x-velocity vx at any
given point equals the slope of the tangent
to the x-t curve at that point.
40
ChApTer 2 Motion Along a Straight Line
2.8 (a) The x-t graph of the motion of a particular particle. (b) A motion diagram showing the position and velocity
of the particle at each of the times labeled on the x-t graph.
(a) x-t graph
Slope zero: vx = 0
x
C
(b) Particle’s motion
D
E
B
0
Slope negative:
vx 6 0
t
tA = 0
v
v
tB
v = 0
tC
0
v
Slope positive:
vx 7 0
0
v
tE
0
The particle is at x 6 0 and moving
in +x-direction.
x From tA to tB it speeds up, ...
0
tD
A
x
0
x
... and from tB to tC it slows down,
then halts momentarily at tC.
x From tC to tD it speeds up in
-x-direction, ...
x ... and from tD to tE it slows down
in -x-direction.
• On an x-t graph, the slope of the tangent at any point equals the particle’s velocity at that point.
• The steeper the slope (positive or negative), the greater the particle’s speed in the positive or negative x-direction.
Figure 2.8 depicts the motion of a particle in two ways: as (a) an x-t graph and
(b) a motion diagram that shows the particle’s position at various instants (like
frames from a video of the particle’s motion) as well as arrows to represent the
particle’s velocity at each instant. We will use both x-t graphs and motion diagrams in this chapter to represent motion. You will find it helpful to draw both an
x-t graph and a motion diagram when you solve any problem involving motion.
2.9 An x-t graph for a particle.
Q
x
P
R
t
O
S
test your unDerstanDing of section 2.2 Figure 2.9 is an x-t graph of
the motion of a particle. (a) Rank the values of the particle’s x-velocity vx at points P, Q, R,
and S from most positive to most negative. (b) At which points is vx positive? (c) At
which points is vx negative? (d) At which points is vx zero? (e) Rank the values of the
particle’s speed at points P, Q, R, and S from fastest to slowest. ❙
2.3 average anD instantaneous
acceLeration
Just as velocity describes the rate of change of position with time, acceleration
describes the rate of change of velocity with time. Like velocity, acceleration is a
vector quantity. When the motion is along a straight line, its only nonzero component is along that line. As we’ll see, acceleration in straight-line motion can
refer to either speeding up or slowing down.
average acceleration
Let’s consider again a particle moving along the x-axis. Suppose that at time t1
the particle is at point P1 and has x-component of (instantaneous) velocity v1x ,
and at a later time t2 it is at point P2 and has x-component of velocity v2x . So
the x-component of velocity changes by an amount ∆vx = v2x - v1x during the
time interval ∆t = t2 - t1 . As the particle moves from P1 to P2 , its average
acceleration is a vector quantity whose x-component aav@x (called the average
x-acceleration) equals ∆vx , the change in the x-component of velocity, divided
by the time interval ∆t:
Average x-acceleration of
a particle in straight-line
motion during time
interval from t1 to t2
Change in x-component of the particle’s velocity
aav-x =
Time interval
v2x - v1x
∆vx
=
∆t
t2 - t1
Final x-velocity
minus initial
x-velocity
Final time minus initial time
(2.4)
2.3 average and Instantaneous acceleration
41
For straight-line motion along the x-axis we will often call aav@x simply the average acceleration. (We’ll encounter the other components of the average acceleration vector in Chapter 3.)
If we express velocity in meters per second and time in seconds, then average
acceleration is in meters per second per second. This is usually written as m>s2
and is read “meters per second squared.”
Caution don’t confuse velocity and acceleration Velocity describes how a body’s
position changes with time; it tells us how fast and in what direction the body moves.
Acceleration describes how the velocity changes with time; it tells us how the speed and
direction of motion change. To see the difference, imagine you are riding along with the
moving body. If the body accelerates forward and gains speed, you feel pushed backward
in your seat; if it accelerates backward and loses speed, you feel pushed forward. If the
velocity is constant and there’s no acceleration, you feel neither sensation. (We’ll explain
these sensations in Chapter 4.) ❙
Solution
ExAmplE 2.2 AvErAgE AccElErAtion
An astronaut has left an orbiting spacecraft to test a new personal
maneuvering unit. As she moves along a straight line, her partner on the spacecraft measures her velocity every 2.0 s, starting at
time t = 1.0 s:
t
vx
t
vx
1.0 s
3.0 s
0.8 m>s
9.0 s
- 0.4 m>s
1.2 m>s
11.0 s
5.0 s
7.0 s
- 1.0 m>s
1.6 m>s
13.0 s
- 1.6 m>s
1.2 m>s
15.0 s
- 0.8 m>s
Find the average x-acceleration, and state whether the speed of
the astronaut increases or decreases over each of these 2.0-s time
intervals: (a) t1 = 1.0 s to t2 = 3.0 s; (b) t1 = 5.0 s to t2 = 7.0 s;
(c) t1 = 9.0 s to t2 = 11.0 s; (d) t1 = 13.0 s to t2 = 15.0 s .
2.10 Our graphs of x-velocity versus time (top) and average
x-acceleration versus time (bottom) for the astronaut.
(b)
(a)
(c)
(d)
The slope of the line connecting each
pair of points on the vx-t graph ...
... equals the average x-acceleration
between those points.
Solution
idEntify and SEt up: We’ll use Eq. (2.4) to determine the average acceleration aav@x from the change in velocity over each
time interval. To find the changes in speed, we’ll use the idea that
speed v is the magnitude of the instantaneous velocity vx .
The upper part of Fig. 2.10 is our graph of the x-velocity as
a function of time. On this vx@t graph, the slope of the line connecting the endpoints of each interval is the average x-acceleration
aav@x = ∆vx >∆t for that interval. The four slopes (and thus the
signs of the average accelerations) are, from left to right, positive,
negative, negative, and positive. The third and fourth slopes (and
thus the average accelerations themselves) have greater magnitude
than the first and second.
ExECutE: Using Eq. (2.4), we find:
(a) aav@x = 11.2 m>s - 0.8 m>s2>13.0 s - 1.0 s2 = 0.2 m>s2.
The speed (magnitude of instantaneous x-velocity) increases from
0.8 m>s to 1.2 m>s.
(b) aav@x = 11.2 m>s - 1.6 m>s2>17.0 s - 5.0 s2 = - 0.2 m>s2.
The speed decreases from 1.6 m>s to 1.2 m>s.
(c) aav@x = 3- 1.0 m>s - 1- 0.4 m>s24>111.0 s - 9.0 s2 =
- 0.3 m>s2. The speed increases from 0.4 m>s to 1.0 m>s.
(d) aav@x = 3-0.8 m>s - 1-1.6 m>s24>115.0 s - 13.0 s2 =
0.4 m>s2. The speed decreases from 1.6 m>s to 0.8 m>s.
In the lower part of Fig. 2.10, we graph the values of aav@x.
EvaluatE: The signs and relative magnitudes of the average accelerations agree with our qualitative predictions. Notice that when
the average x-acceleration has the same algebraic sign as the initial
velocity, as in intervals (a) and (c), the astronaut goes faster. When
aav@x has the opposite algebraic sign from the initial velocity, as in
intervals (b) and (d), she slows down. Thus positive x-acceleration
means speeding up if the x-velocity is positive [interval (a)] but
slowing down if the x-velocity is negative [interval (d)]. Similarly,
negative x-acceleration means speeding up if the x-velocity is
negative [interval (c)] but slowing down if the x-velocity is positive
[interval (b)].
42
ChApTer 2 Motion Along a Straight Line
2.11 A Grand Prix car at two points on the straightaway.
Speed v1
x-velocity v1x
O
Speed v2
x-velocity v2 x
P1
x
P2
instantaneous acceleration
We can now define instantaneous acceleration by following the same procedure that we used to define instantaneous velocity. Suppose a race car driver is
driving along a straightaway as shown in Fig. 2.11. To define the instantaneous
acceleration at point P1 , we take point P2 in Fig. 2.11 to be closer and closer to
P1 so that the average acceleration is computed over shorter and shorter time
intervals. Thus
The instantaneous
x-acceleration of a particle
in straight-line motion ...
dv
∆vx
= x
∆tS0 ∆t
dt
ax = lim
... equals the limit of the particle’s average
x-acceleration as the time interval approaches zero ...
(2.5)
... and equals the instantaneous rate
of change of the particle’s x-velocity.
In Eq. (2.5) ax is the x-component of the acceleration vector, which we call the
instantaneous x-acceleration; in straight-line motion, all other components of
this vector are zero. From now on, when we use the term “acceleration,” we will
always mean instantaneous acceleration, not average acceleration.
Solution
ExAmplE 2.3 AvErAgE And insTAnTAnEous AccElErATions
Suppose the x-velocity vx of the car in Fig. 2.11 at any time t is
given by the equation
The change in x-velocity ∆vx between t1 = 1.0 s and t2 = 3.0 s is
vx = 60 m>s + 10.50 m>s32t 2
(b) The average x-acceleration during this time interval of
duration t2 - t1 = 2.0 s is
(a) Find the change in x-velocity of the car in the time interval
t1 = 1.0 s to t2 = 3.0 s. (b) Find the average x-acceleration in this
time interval. (c) Find the instantaneous x-acceleration at time
t1 = 1.0 s by taking ∆t to be first 0.1 s, then 0.01 s, then 0.001 s.
(d) Derive an expression for the instantaneous x-acceleration as a
function of time, and use it to find ax at t = 1.0 s and t = 3.0 s.
soLution
iDentify and set up: This example is analogous to Example 2.1
in Section 2.2. In that example we found the average x-velocity
from the change in position over shorter and shorter time intervals,
and we obtained an expression for the instantaneous x-velocity by
differentiating the position as a function of time. In this example
we have an exact parallel. Using Eq. (2.4), we’ll find the average
x-acceleration from the change in x-velocity over a time interval.
Likewise, using Eq. (2.5), we’ll obtain an expression for the instantaneous x-acceleration by differentiating the x-velocity as a function
of time.
execute: (a) Before we can apply Eq. (2.4), we must find the
x-velocity at each time from the given equation. At t1 = 1.0 s and
t2 = 3.0 s, the velocities are
3
2
v1x = 60 m>s + 10.50 m>s 211.0 s2 = 60.5 m>s
v2x = 60 m>s + 10.50 m>s3213.0 s22 = 64.5 m>s
∆vx = v2x - v1x = 64.5 m>s - 60.5 m>s = 4.0 m>s
aav@x =
v2x - v1x
4.0 m>s
= 2.0 m>s2
=
t2 - t1
2.0 s
During this time interval the x-velocity and average x-acceleration
have the same algebraic sign (in this case, positive), and the car
speeds up.
(c) When ∆t = 0.1 s, we have t2 = 1.1 s. Proceeding as before, we find
v2x = 60 m>s + 10.50 m>s3211.1 s22 = 60.605 m>s
∆vx = 0.105 m>s
aav@x =
∆vx
0.105 m>s
=
= 1.05 m>s2
∆t
0.1 s
You should follow this pattern to calculate aav@x for ∆t = 0.01 s
and ∆t = 0.001 s; the results are aav@x = 1.005 m>s2 and
aav@x = 1.0005 m>s2, respectively. As ∆t gets smaller, the average x-acceleration gets closer to 1.0 m>s2, so the instantaneous
x-acceleration at t = 1.0 s is 1.0 m>s2.
(d) By Eq. (2.5) the instantaneous x-acceleration is ax = dvx>dt.
The derivative of a constant is zero and the derivative of t 2 is 2t, so
ax =
dvx
d
=
360 m>s + 10.50 m>s32t 24
dt
dt
= 10.50 m>s3212t2 = 11.0 m>s32t
2.3 average and Instantaneous acceleration
When t = 1.0 s,
3
43
EValuatE: Neither of the values we found in part (d) is equal
to the average x-acceleration found in part (b). That’s because
the car’s instantaneous x-acceleration varies with time. The
rate of change of acceleration with time is sometimes called
the “jerk.”
2
ax = 11.0 m>s 211.0 s2 = 1.0 m>s
When t = 3.0 s,
ax = 11.0 m>s3213.0 s2 = 3.0 m>s2
finding acceleration on a vx -t Graph or an x-t Graph
In Section 2.2 we interpreted average and instantaneous x-velocity in terms of
the slope of a graph of position versus time. In the same way, we can interpret
average and instantaneous x-acceleration by using a graph of instantaneous velocity vx versus time t—that is, a vx@t graph (Fig. 2.12). Points p1 and p2 on the
graph correspond to points P1 and P2 in Fig. 2.11. The average x-acceleration
aav@x = ∆vx >∆t during this interval is the slope of the line p1 p2 .
As point P2 in Fig. 2.11 approaches point P1 , point p2 in the vx@t graph of
Fig. 2.12 approaches point p1 , and the slope of the line p1 p2 approaches the slope
of the line tangent to the curve at point p1 . Thus, on a graph of x-velocity as
a function of time, the instantaneous x-acceleration at any point is equal to
the slope of the tangent to the curve at that point. Tangents drawn at different
points along the curve in Fig. 2.12 have different slopes, so the instantaneous
x-acceleration varies with time.
?
Caution Signs of x-acceleration and x-velocity The algebraic sign of the
x-acceleration does not tell you whether a body is speeding up or slowing down.
You must compare the signs of the x-velocity and the x-acceleration. When vx and ax
have the same sign, the body is speeding up. If both are positive, the body is moving in
the positive direction with increasing speed. If both are negative, the body is moving in
the negative direction with an x-velocity that is becoming more negative, and again the
speed is increasing. When vx and ax have opposite signs, the body is slowing down. If vx is
positive and ax is negative, the body is moving in the positive direction with decreasing
speed; if vx is negative and ax is positive, the body is moving in the negative direction
with an x-velocity that is becoming less negative, and again the body is slowing down.
Table 2.3 summarizes these rules, and Fig. 2.13 (next page) illustrates some of them. ❙
The term “deceleration” is sometimes used for a decrease in speed. Because it
may mean positive or negative ax , depending on the sign of vx , we avoid this term.
We can also learn about the acceleration of a body from a graph of its position
versus time. Because ax = dvx >dt and vx = dx>dt, we can write
dvx
d dx
d 2x
ax =
=
a b = 2
dt
dt dt
dt
vx
For a displacement along the x-axis, an object’s average x-acceleration
equals the slope of a line connecting the corresponding points on a
graph of x-velocity (vx) versus time (t).
p2
v2x
e
op
O
p1
Sl
=
e
av
c
-a
ex
g
ra
le
ce
n
tio
ra
v1x
(2.6)
∆vx = v2x - v1x
Slope of tangent to vx-t curve at a given point
= instantaneous x-acceleration at that point.
∆t = t2 - t1
t1
t2
t
Rules for the Sign of
Table 2.3 x-acceleration
If x-velocity is:
. . . x-acceleration is:
Positive & increasing
(getting more positive)
Positive: Particle
is moving in
+ x-direction &
speeding up
Positive & decreasing
(getting less positive)
Negative: Particle
is moving in
+ x-direction &
slowing down
Negative & increasing
(getting less negative)
Positive: Particle
is moving in
- x-direction &
slowing down
Negative & decreasing Negative: Particle
(getting more negative) is moving in
- x-direction &
speeding up
Note: These rules apply to both the average
x-acceleration aav@x and the instantaneous
x-acceleration ax.
2.12 A vx@t graph of the motion in
Fig. 2.11.
44
ChApTer 2 Motion Along a Straight Line
2.13 (a) The vx@t graph of the motion of a different particle from that shown in Fig. 2.8. (b) A motion diagram
showing the position, velocity, and acceleration of the particle at each of the times labeled on the vx@t graph.
(a) vx-t graph
vx
Slope zero: ax = 0
(b) Particle’s motion
a
v
tA = 0
C
0
A
0
x
Particle is at x 6 0, moving in -x-direction (vx 6 0),
and slowing down (vx and ax have opposite signs).
x
Particle is at x 6 0, instantaneously at rest (vx = 0), and
about to move in +x-direction (ax 7 0).
x
Particle is at x 7 0, moving in +x-direction (vx 7 0);
its speed is instantaneously not changing (ax = 0).
a
B
D
Slope positive:
ax 7 0
tB
t
v = 0
0
a = 0
tC
E
v
0
a
Slope negative:
ax 6 0
tD
v = 0
0
Particle is at x 7 0, instantaneously at rest (vx = 0), and
x about to move in -x-direction (a 6 0).
x
a
Particle is at x 7 0, moving in -x-direction (vx 6 0),
v
x
and speeding up (vx and ax have the same sign).
0
• On a vx-t graph, the slope of the tangent at any point equals the particle’s acceleration at that point.
• The steeper the slope (positive or negative), the greater the particle’s acceleration in the positive or negative x-direction.
tE
That is, ax is the second derivative of x with respect to t. The second derivative
of any function is directly related to the concavity or curvature of the graph of
that function (Fig. 2.14). At a point where the x-t graph is concave up (curved
upward), such as point A or E in Fig. 2.14a, the x-acceleration is positive and vx is
increasing. At a point where the x-t graph is concave down (curved downward),
such as point C in Fig. 2.14a, the x-acceleration is negative and vx is decreasing.
At a point where the x-t graph has no curvature, such as the inflection points B
and D in Fig. 2.14a, the x-acceleration is zero and the velocity is not changing.
Examining the curvature of an x-t graph is an easy way to identify the sign of
acceleration. This technique is less helpful for determining numerical values of
acceleration because the curvature of a graph is hard to measure accurately.
test your unDerstanDing of section 2.3 Look again at the x-t graph
in Fig. 2.9 at the end of Section 2.2. (a) At which of the points P, Q, R, and S is the
x-acceleration ax positive? (b) At which points is the x-acceleration negative? (c) At
which points does the x-acceleration appear to be zero? (d) At each point state whether
the velocity is increasing, decreasing, or not changing. ❙
2.14 (a) The same x-t graph as shown in Fig. 2.8a. (b) A motion diagram showing the position, velocity, and
acceleration of the particle at each of the times labeled on the x-t graph.
(b) Particle’s motion
(a) x-t graph
Slope zero: vx = 0
Curvature downward: ax 6 0
x
C
0
B
a
Slope negative:
tA = 0
vx 6 0
Curvature upward:
D ax 7 0
tB
E
t
tC
Slope negative: v 6 0
v
0
a = 0
A
Slope positive: vx 7 0
Curvature upward: ax 7 0
v
0
a
v = 0
0
x
Curvature zero: ax = 0
Slope positive: vx 7 0
Curvature zero: ax = 0
x
v a = 0
tD
0
x
x
x
a
tE
v
0
x
Particle is at x 6 0, moving in +x-direction
(vx 7 0) and speeding up (vx and ax have the
same sign).
Particle is at x = 0, moving in +x-direction
(vx 7 0); speed is instantaneously not
changing (ax = 0).
Particle is at x 7 0, instantaneously at rest
(vx = 0) and about to move in -x-direction
(ax 6 0).
Particle is at x 7 0, moving in -x-direction
(vx 6 0); speed is instantaneously not
changing (ax = 0).
Particle is at x 7 0, moving in -x-direction
(vx 6 0) and slowing down (vx and ax have
opposite signs).
• On an x-t graph, the curvature at any point tells you the particle’s acceleration at that point.
• The greater the curvature (positive or negative), the greater the particle’s acceleration in the positive or negative x-direction.
45
2.4 Motion with Constant Acceleration
2.4 motion with constant acceLeration
The simplest kind of accelerated motion is straight-line motion with constant
acceleration. In this case the velocity changes at the same rate throughout the
motion. As an example, a falling body has a constant acceleration if the effects
of the air are not important. The same is true for a body sliding on an incline or
along a rough horizontal surface, or for an airplane being catapulted from the
deck of an aircraft carrier.
Figure 2.15 is a motion diagram showing the position, velocity, and acceleration of a particle moving with constant acceleration. Figures 2.16 and 2.17 depict
this same motion in the form of graphs. Since the x-acceleration is constant, the
ax@t graph (graph of x-acceleration versus time) in Figure 2.16 is a horizontal line.
The graph of x-velocity versus time, or vx@t graph, has a constant slope because
the acceleration is constant, so this graph is a straight line (Fig. 2.17).
When the x-acceleration ax is constant, the average x-acceleration aav@x for any
time interval is the same as ax . This makes it easy to derive equations for the
position x and the x-velocity vx as functions of time. To find an expression for vx ,
we first replace aav@x in Eq. (2.4) by ax :
v2x - v1x
ax =
t2 - t1
2.15 A motion diagram for a particle
moving in a straight line in the positive
x-direction with constant positive
x-acceleration ax .
If a particle moves in a
straight line with constant
x-acceleration ax ...
a
v
0
t = 0
x
... the x-velocity changes
by equal amounts in equal
time intervals.
a
v
t = ∆t
0
x
a
v
t = 2∆t
0
x
a
v
t = 3∆t
0
x
a
v
t = 4∆t
0
x
However, the position changes by different
amounts in equal time intervals because the
velocity is changing.
(2.7)
Now we let t1 = 0 and let t2 be any later time t. We use the symbol v0x for the
initial x-velocity at time t = 0; the x-velocity at the later time t is vx . Then Eq. (2.7)
becomes
ax =
vx - v0x
t - 0
2.16 An acceleration-time 1ax@t2 graph of
straight-line motion with constant positive
x-acceleration ax .
or
x-velocity at time t of
x-velocity of the particle at time 0
a particle with
vx = v0x + ax t
constant x-acceleration
Constant x-acceleration of the particle
Time
ax
(2.8)
ax
In Eq. (2.8) the term ax t is the product of the constant rate of change of
x-velocity, ax , and the time interval t. Therefore it equals the total change in
x-velocity from t = 0 to time t. The x-velocity vx at any time t then equals the
initial x-velocity v0x (at t = 0) plus the change in x-velocity ax t (Fig. 2.17).
Equation (2.8) also says that the change in x-velocity vx - v0x of the particle
between t = 0 and any later time t equals the area under the ax@t graph between
those two times. You can verify this from Fig. 2.16: Under this graph is a rectangle of vertical side ax, horizontal side t, and area axt. From Eq. (2.8) the area
axt is indeed equal to the change in velocity vx - v0x . In Section 2.6 we’ll show
that even if the x-acceleration is not constant, the change in x-velocity during a
time interval is still equal to the area under the ax@t curve, although then Eq. (2.8)
does not apply.
Next we’ll derive an equation for the position x as a function of time when the
x-acceleration is constant. To do this, we use two different expressions for the average x-velocity vav@x during the interval from t = 0 to any later time t. The first
expression comes from the definition of vav@x , Eq. (2.2), which is true whether
or not the acceleration is constant. We call the position at time t = 0 the initial
position, denoted by x0 . The position at time t is simply x. Thus for the time interval ∆t = t - 0 the displacement is ∆x = x - x0 , and Eq. (2.2) gives
vav@x =
x - x0
t
Constant x-acceleration: ax-t graph
is a horizontal line (slope = 0).
Area under ax-t graph = vx - v0x
= change in x-velocity from time 0 to time t.
2.17 A velocity-time 1vx@t2 graph of
straight-line motion with constant positive
x-acceleration ax . The initial x-velocity v0x
is also positive in this case.
vx
Constant
x-acceleration:
vx-t graph is a
straight line.
vx
During time
interval t, the
x-velocity changes
by vx - v0x = axt.
tion
era
el
acc
v0x
pe
Slo
=x
ax t
vx
v0x
(2.9)
To find a second expression for vav@x, note that the x-velocity changes at a
constant rate if the x-acceleration is constant. In this case the average x-velocity
t
t
O
O
t
t
Total area under vx -t graph = x - x0
= change in x-coordinate from time 0 to time t.
46
Chapter 2 Motion along a Straight Line
for the time interval from 0 to t is simply the average of the x-velocities at the
beginning and end of the interval:
PhET: Forces in 1 Dimension
BIO application Testing Humans at
High accelerations In experiments
carried out by the U.S. Air Force in the 1940s
and 1950s, humans riding a rocket sled could
withstand accelerations as great as 440 m>s2.
The first three photos in this sequence show
Air Force physician John Stapp speeding up
from rest to 188 m>s (678 km>h = 421 mi>h)
in just 5 s. Photos 4–6 show the even greater
magnitude of acceleration as the rocket sled
braked to a halt.
vav@x = 12 (v0x + vx)
(constant x-acceleration only)
(2.10)
[Equation (2.10) is not true if the x-acceleration varies during the time interval.]
We also know that with constant x-acceleration, the x-velocity vx at any time t is
given by Eq. (2.8). Substituting that expression for vx into Eq. (2.10), we find
vav@x = 12 1v0x + v0x + ax t2 (constant
x-acceleration only)
= v0x + 12 ax t
(2.11)
Finally, we set Eqs. (2.9) and (2.11) equal to each other and simplify:
v0x + 12 ax t =
x - x0
t
or
Position of the particle at time 0
Time
Position at time t of a
1
2
particle with constant
x = x0 + v0x t + 2 ax t
x-acceleration
x-velocity of the particle at time 0
Constant x-acceleration of the particle
PhET: The Moving Man
2.18 (a) Straight-line motion with constant acceleration. (b) A position-time 1x@t2
graph for this motion (the same motion as
is shown in Figs. 2.15, 2.16, and 2.17). For
this motion the initial position x0 , the
initial velocity v0x , and the acceleration ax
are all positive.
(2.12)
Equation (2.12) tells us that the particle’s position at time t is the sum of three
terms: its initial position at t = 0, x0, plus the displacement v0x t it would have if
its x-velocity remained equal to its initial value, plus an additional displacement
1
2
2 ax t caused by the change in x-velocity.
A graph of Eq. (2.12)—that is, an x-t graph for motion with constant
x-acceleration (Fig. 2.18a)—is always a parabola. Figure 2.18b shows such a
graph. The curve intercepts the vertical axis (x-axis) at x0 , the position at t = 0.
The slope of the tangent at t = 0 equals v0x , the initial x-velocity, and the slope
of the tangent at any time t equals the x-velocity vx at that time. The slope and
x-velocity are continuously increasing, so the x-acceleration ax is positive and
the graph in Fig. 2.18b is concave up (it curves upward). If ax is negative, the x-t
graph is a parabola that is concave down (has a downward curvature).
If there is zero x-acceleration, the x-t graph is a straight line; if there is a constant
x-acceleration, the additional 12 ax t 2 term in Eq. (2.12) for x as a function of t curves
the graph into a parabola (Fig. 2.19a). Similarly, if there is zero x-acceleration,
the vx@t graph is a horizontal line (the x-velocity is constant). Adding a constant
x-acceleration in Eq. (2.8) gives a slope to the graph (Fig. 2.19b).
Here’s another way to derive Eq. (2.12). Just as the change in x-velocity of the
particle equals the area under the ax@t graph, the displacement (change in position)
equals the area under the vx@t graph. So the displacement x - x0 of the particle
between t = 0 and any later time t equals the area under the vx@t graph between
(a) A race car moves in the x-direction
with constant acceleration.
(b) The x-t graph
x
x
vx = v0x + ax t
x
Slope = vx
x
During time interval t,
the x-velocity changes
by vx - v0x = ax t.
Constant x-acceleration:
x-t graph is a parabola.
v0x
x0
x0
O
O
Slope = v0x
t
t
47
2.4 Motion with Constant Acceleration
vx
Graph for constant x-acceleration:
vx = v0x + ax t
The effect of
x-acceleration:
1
2
2 ax t
Graph for zero
x-acceleration:
x = x0 + v0x t
t
x0
O
2.19 (a) How a constant x-acceleration
affects a particle’s (a) x-t graph and
(b) vx@t graph.
(b) The vx-t graph for the same particle
(a) An x-t graph for a particle moving with
positive constant x-acceleration
Graph for constant x-acceleration:
x
x = x0 + v0x t + 12 ax t 2
The added velocity
due to x-acceleration:
ax t
v0x
Graph for zero x-acceleration:
vx = v0x
t
O
those times. In Fig. 2.17 we divide the area under the graph into a dark-colored
rectangle (vertical side v0x, horizontal side t, and area v0xt) and a light-colored right
triangle (vertical side axt, horizontal side t, and area 12 1axt21t2 = 21 axt 2). The
total area under the vx@t graph is x - x0 = v0x t + 12 ax t 2, in accord with Eq. (2.12).
The displacement during a time interval is always equal to the area under the
vx@t curve. This is true even if the acceleration is not constant, although in that
case Eq. (2.12) does not apply. (We’ll show this in Section 2.6.)
It’s often useful to have a relationship for position, x-velocity, and (constant)
x-acceleration that does not involve time. To obtain this, we first solve Eq. (2.8)
for t and then substitute the resulting expression into Eq. (2.12):
t =
vx - v0x
ax
x = x0 + v0x a
vx - v0x
vx - v0x 2
b + 12 ax a
b
ax
ax
We transfer the term x0 to the left side, multiply through by 2ax, and simplify:
Finally,
2ax 1x - x02 = 2v0x vx - 2v 0x 2 + v x2 - 2v0x vx + v 0x 2
x-velocity of the particle at time 0
x-velocity at time t of
2
2
a particle with
vx = v0x + 2ax1x - x02
constant x-acceleration
Constant x-acceleration
Position of the
Position of the
of the particle
particle at time t
particle at time 0
(2.13)
We can get one more useful relationship by equating the two expressions for
vav@x , Eqs. (2.9) and (2.10), and multiplying through by t:
Position at time t of
a particle with
constant x-acceleration
Position of the particle at time 0
x - x0 =
x-velocity of the particle at time 0
1
21v0x
+ vx2t
Time
(2.14)
x-velocity of the particle at time t
Note that Eq. (2.14) does not contain the x-acceleration ax . This equation can be
handy when ax is constant but its value is unknown.
Equations (2.8), (2.12), (2.13), and (2.14) are the equations of motion with
constant acceleration (Table 2.4). By using these equations, we can solve any
problem involving straight-line motion of a particle with constant acceleration.
For the particular case of motion with constant x-acceleration depicted in
Fig. 2.15 and graphed in Figs. 2.16, 2.17, and 2.18, the values of x0 , v0x , and ax
are all positive. We recommend that you redraw these figures for cases in which
one, two, or all three of these quantities are negative.
equations of motion with
Table 2.4 constant acceleration
Includes
Quantities
Equation
vx = v0x + ax t
(2.8)
x = x0 + v0x t + 12 ax t 2
(2.12) t
v x2
2
= v 0x + 2ax 1x - x02 (2.13)
x - x0 = 12 1v0x + vx2t
vx ax
t
(2.14) t
x
ax
x vx ax
x vx
48
ChApTer 2 Motion Along a Straight Line
problEm-solving sTrATEgy 2.1
moTion wiTH consTAnT AccElErATion
iDentify the relevant concepts: In most straight-line motion problems, you can use the constant-acceleration equations (2.8), (2.12),
(2.13), and (2.14). If you encounter a situation in which the acceleration isn’t constant, you’ll need a different approach (see Section 2.6).
set up the problem using the following steps:
1. Read the problem carefully. Make a motion diagram showing
the location of the particle at the times of interest. Decide
where to place the origin of coordinates and which axis direction is positive. It’s often helpful to place the particle at the origin at time t = 0; then x0 = 0. Your choice of the positive axis
direction automatically determines the positive directions for
x-velocity and x-acceleration. If x is positive to the right of the
origin, then vx and ax are also positive toward the right.
2. Identify the physical quantities (times, positions, velocities, and
accelerations) that appear in Eqs. (2.8), (2.12), (2.13), and (2.14)
and assign them appropriate symbols: x, x0, vx, v0x, and ax, or
symbols related to those. Translate the prose into physics: “When
does the particle arrive at its highest point” means “What is the
value of t when x has its maximum value?” In Example 2.4,
“Where is he when his speed is 25 m>s?” means “What is the
value of x when vx = 25 m>s?” Be alert for implicit information.
For example, “A car sits at a stop light” usually means v0x = 0.
3. List the quantities such as x, x0, vx, v0x, ax, and t. Some of them
will be known and some will be unknown. Write down the values
of the known quantities, and decide which of the unknowns are
the target variables. Make note of the absence of any of the quantities that appear in the four constant-acceleration equations.
4. Use Table 2.4 to identify the applicable equations. (These are
often the equations that don’t include any of the absent quantities
that you identified in step 3.) Usually you’ll find a single equation
that contains only one of the target variables. Sometimes you must
find two equations, each containing the same two unknowns.
5. Sketch graphs corresponding to the applicable equations. The
vx@t graph of Eq. (2.8) is a straight line with slope ax. The x@t
graph of Eq. (2.12) is a parabola that’s concave up if ax is positive and concave down if ax is negative.
6. On the basis of your experience with such problems, and taking
account of what your sketched graphs tell you, make any qualitative and quantitative predictions you can about the solution.
execute the solution: If a single equation applies, solve it for the
target variable, using symbols only; then substitute the known values and calculate the value of the target variable. If you have two
equations in two unknowns, solve them simultaneously for the target variables.
evaLuate your answer: Take a hard look at your results to see
whether they make sense. Are they within the general range of
values that you expected?
Solution
ExAmplE 2.4 consTAnT-AccElErATion cAlculATions
A motorcyclist heading east through a small town accelerates
at a constant 4.0 m>s2 after he leaves the city limits (Fig. 2.20).
At time t = 0 he is 5.0 m east of the city-limits signpost while
he moves east at 15 m>s. (a) Find his position and velocity at
t = 2.0 s. (b) Where is he when his speed is 25 m>s?
iDentify and set up: The x-acceleration is constant, so we can
use the constant-acceleration equations. We take the signpost as
the origin of coordinates 1x = 02 and choose the positive x-axis
to point east (see Fig. 2.20, which is also a motion diagram). The
known variables are the initial position and velocity, x0 = 5.0 m
and v0x = 15 m>s, and the acceleration, ax = 4.0 m>s2. The unknown target variables in part (a) are the values of the position x
and the x-velocity vx at t = 2.0 s; the target variable in part (b) is
the value of x when vx = 25 m>s.
execute: (a) Since we know the values of x0, v0x, and ax, Table 2.4
tells us that we can find the position x at t = 2.0 s by using
= 5.0 m + 115 m>s212.0 s2 +
1
2
= 43 m
ax = 4.0 m>s2
v0x = 15 m>s
x0 = 5.0 m
t = 0
vx = ?
x = ?
t = 2.0 s
x (east)
14.0 m>s2212.0 s22
= 15 m>s + 14.0 m>s2212.0 s2 = 23 m>s
(b) We want to find the value of x when vx = 25 m>s, but
we don’t know the time when the motorcycle has this velocity.
Table 2.4 tells us that we should use Eq. (2.13), which involves
x, vx, and ax but does not involve t:
v x2 = v 0x2 + 2ax1x - x02
Solving for x and substituting the known values, we find
x = x0 +
v x2 - v 0x 2
2ax
= 5.0 m +
2.20 A motorcyclist traveling with constant acceleration.
O
x = x0 + v0x t + 12 ax t 2
vx = v0x + ax t
soLution
OSAGE
Eq. (2.12) and the x-velocity vx at this time by using Eq. (2.8):
125 m>s22 - 115 m>s22
214.0 m>s22
= 55 m
evaLuate: You can check the result in part (b) by first using
Eq. (2.8), vx = v0x + axt, to find the time at which vx = 25 m>s,
which turns out to be t = 2.5 s. You can then use Eq. (2.12),
x = x0 + v0xt + 12 axt 2, to solve for x. You should find x = 55 m,
the same answer as above. That’s the long way to solve the problem,
though. The method we used in part (b) is much more efficient.
49
2.4 Motion with Constant Acceleration
Solution
ExAmplE 2.5 Two bodiEs wiTH diffErEnT AccElErATions
A motorist traveling at a constant 15 m>s (about 34 mi>h) passes a
school crossing where the speed limit is 10 m>s (about 22 mi>h).
Just as the motorist passes the school-crossing sign, a police officer on a motorcycle stopped there starts in pursuit with constant
acceleration 3.0 m>s2 (Fig. 2.21a). (a) How much time elapses
before the officer passes the motorist? At that time, (b) what is the
officer’s speed and (c) how far has each vehicle traveled?
soLution
iDentify and set up: Both the officer and the motorist move
with constant acceleration (equal to zero for the motorist), so we
can use the constant-acceleration formulas. We take the origin at
the sign, so x0 = 0 for both, and we take the positive direction to
the right. Let xP and xM represent the positions of the police officer and the motorist at any time. Their initial velocities are vP0x = 0
and vM0x = 15 m>s, and their accelerations are aPx = 3.0 m>s2
and aMx = 0. Our target variable in part (a) is the time when the
officer and motorist are at the same position x; Table 2.4 tells us
that Eq. (2.12) is useful for this part. In part (b) we’ll use Eq. (2.8) to
find the officer’s speed v (the magnitude of her velocity) at the
time found in part (a). In part (c) we’ll use Eq. (2.12) again to find
the position of either vehicle at this same time.
Figure 2.21b shows an x@t graph for both vehicles. The straight
line represents the motorist’s motion, xM = xM0 + vM0x t = vM0x t.
The graph for the officer’s motion is the right half of a parabola
with upward curvature:
xP = xP0 + vP0x t + 12 aPx t 2 = 12 aPx t 2
A good sketch shows that the officer and motorist are at the same
position 1xP = xM2 at about t = 10 s, at which time both have
traveled about 150 m from the sign.
execute: (a) To find the value of the time t at which the motorist
and police officer are at the same position, we set xP = xM by equating the expressions above and solving that equation for t:
Both vehicles have the same x-coordinate at two times, as Fig. 2.21b
indicates. At t = 0 the motorist passes the officer; at t = 10 s the
officer passes the motorist.
(b) We want the magnitude of the officer’s x-velocity vPx at the
time t found in part (a). Substituting the values of vP0x and aPx into
Eq. (2.8) along with t = 10 s from part (a), we find
vPx = vP0x + aPx t = 0 + 13.0 m>s22110 s2 = 30 m>s
The officer’s speed is the absolute value of this, which is also
30 m>s.
(c) In 10 s the motorist travels a distance
xM = vM0x t = 115 m>s2110 s2 = 150 m
and the officer travels
xP = 12 aPx t 2 = 1213.0 m>s22110 s22 = 150 m
This verifies that they have gone equal distances after 10 s.
evaLuate: Our results in parts (a) and (c) agree with our estimates
from our sketch. Note that when the officer passes the motorist, they do not have the same velocity: The motorist is moving
at 15 m>s and the officer is moving at 30 m>s. You can also see
this from Fig. 2.21b. Where the two x@t curves cross, their slopes
(equal to the values of vx for the two vehicles) are different.
Is it just coincidence that when the two vehicles are at the
same position, the officer is going twice the speed of the motorist?
Equation (2.14), x - x0 = 12 1v0x + vx2t, gives the answer. The
motorist has constant velocity, so vM0x = vMx , and the motorist’s
displacement x - x0 in time t is vM0x t. Because vP0x = 0, in the
same time t the officer’s displacement is 12 vPx t. The two vehicles
have the same displacement in the same amount of time, so
vM0xt = 12 vPxt and vPx = 2vM0x—that is, the officer has exactly
twice the motorist’s velocity. This is true no matter what the value
of the officer’s acceleration.
vM0x t = 12 aPx t 2
t = 0
or
t =
2 115 m>s2
2vM0x
=
= 10 s
aPx
3.0 m>s2
2.21 (a) Motion with constant acceleration overtaking motion with constant velocity.
(b) A graph of x versus t for each vehicle.
(b)
The police officer and motorist
meet at the time t where their
x ( m) x-t graphs cross.
(a)
160
SCHOOL
CROSSING
Police officer: initially at rest,
constant x-acceleration
120
Motorist: constant x-velocity
aPx = 3.0 m>s2
O
xP
vM0x = 15 m>s
xM
Motorist
80
40
x
O
Officer
2
4
6
8
10
12
t ( s)
50
Chapter 2 Motion along a Straight Line
TesT Your undersTanding of secTion 2.4 Four possible vx@t graphs are
shown for the two vehicles in Example 2.5. Which graph is correct?
(a)
2.22 Multiflash photo of a freely falling
(b)
vx
ball.
vx
Officer
O
10
O
vx
Motorist
Officer
t (s)
(d)
vx
Motorist
Motorist
The ball is
released
here and
falls freely.
(c)
Officer
t (s)
10
Motorist
O
Officer
t ( s)
10
O
t ( s)
10
❙
2.5 freelY falling Bodies
Motion
The images
of the ball
are recorded
at equal time
intervals.
• The average velocity in each time interval is
proportional to the distance between images.
• This distance continuously increases, so the
ball’s velocity is continuously changing; the
ball is accelerating downward.
PhET: Lunar Lander
daTa SpeakS
free fall
When students were given a problem
about free fall, more than 20% gave an
incorrect answer. Common errors:
●
●
Confusing speed, velocity, and acceleration. Speed can never be negative;
velocity can be positive or negative,
depending on the direction of motion.
In free fall, speed and velocity change
continuously but acceleration (rate
of change of velocity) is constant and
downward.
Not realizing that a freely falling body
that moves upward at a certain speed
past a point will pass that same point at
the same speed as it moves downward
(see Example 2.7).
The most familiar example of motion with (nearly) constant acceleration is
a body falling under the influence of the earth’s gravitational attraction. Such
motion has held the attention of philosophers and scientists since ancient times.
In the fourth century b.c., Aristotle thought (erroneously) that heavy bodies fall
faster than light bodies, in proportion to their weight. Nineteen centuries later,
Galileo (see Section 1.1) argued that a body should fall with a downward acceleration that is constant and independent of its weight.
Experiment shows that if the effects of the air can be ignored, Galileo is right;
all bodies at a particular location fall with the same downward acceleration,
regardless of their size or weight. If in addition the distance of the fall is small
compared with the radius of the earth, and if we ignore small effects due to
the earth’s rotation, the acceleration is constant. The idealized motion that
results under all of these assumptions is called free fall, although it includes
rising as well as falling motion. (In Chapter 3 we will extend the discussion of
free fall to include the motion of projectiles, which move both vertically and
horizontally.)
Figure 2.22 is a photograph of a falling ball made with a stroboscopic light
source that produces a series of short, intense flashes at equal time intervals. As
each flash occurs, an image of the ball at that instant is recorded on the photograph. The increasing spacing between successive images in Fig. 2.22 indicates
that the ball is accelerating downward. Careful measurement shows that the
velocity change is the same in each time interval, so the acceleration of the freely
falling ball is constant.
The constant acceleration of a freely falling body is called the acceleration
due to gravity, and we denote its magnitude with the letter g. We will frequently
use the approximate value of g at or near the earth’s surface:
g = 9.80 m>s2 = 980 cm>s2 = 32.2 ft>s2
(approximate value near the
earth’s surface)
The exact value varies with location, so we will often give the value of g at the
earth’s surface to only two significant figures as 9.8 m>s2. On the moon’s surface,
the acceleration due to gravity is caused by the attractive force of the moon rather
than the earth, and g = 1.6 m>s2. Near the surface of the sun, g = 270 m>s2.
cauTion g is always a positive number Because g is the magnitude of a vector quantity, it is always a positive number. If you take the positive direction to be upward, as we do
in most situations involving free fall, the acceleration is negative (downward) and equal to
-g. Be careful with the sign of g, or you’ll have trouble with free-fall problems. ❙
In the following examples we use the constant-acceleration equations developed in Section 2.4. Review Problem-Solving Strategy 2.1 in that section before
you study the next examples.
2.5 Freely Falling Bodies
Solution
ExAmplE 2.6 A frEEly fAlling coin
A one-euro coin is dropped from the Leaning Tower of Pisa and
falls freely from rest. What are its position and velocity after 1.0 s,
2.0 s, and 3.0 s?
Solution
idEntify and SEt up: “Falls freely” means “falls with constant
acceleration due to gravity,” so we can use the constant-acceleration
equations. The right side of Fig. 2.23 shows our motion diagram
2.23 A coin freely falling from rest.
51
for the coin. The motion is vertical, so we use a vertical coordinate axis and call the coordinate y instead of x. We take the origin O at the starting point and the upward direction as positive.
Both the initial coordinate y0 and initial y-velocity v0y are zero.
The y-acceleration is downward (in the negative y-direction), so
ay = - g = -9.8 m>s2. (Remember that, by definition, g is a positive quantity.) Our target variables are the values of y and vy at
the three given times. To find these, we use Eqs. (2.12) and (2.8)
with x replaced by y. Our choice of the upward direction as positive means that all positions and velocities we calculate will be
negative.
ExECutE: At a time t after the coin is dropped, its position and
y-velocity are
y = y0 + v0y t + 12 ay t 2 = 0 + 0 + 12 1- g2t 2 = 1- 4.9 m>s22t 2
vy = v0y + ay t = 0 + 1-g2t = 1-9.8 m>s22t
When t = 1.0 s, y = 1- 4.9 m>s2211.0 s22 = - 4.9 m and vy =
1- 9.8 m>s2211.0 s2 = - 9.8 m>s; after 1.0 s, the coin is 4.9 m
below the origin (y is negative) and has a downward velocity (vy is
negative) with magnitude 9.8 m>s.
We can find the positions and y-velocities at 2.0 s and 3.0 s in
the same way. The results are y = -20 m and vy = - 20 m>s at
t = 2.0 s, and y = -44 m and vy = -29 m>s at t = 3.0 s .
EvaluatE: All our answers are negative, as we expected. If we
had chosen the positive y-axis to point downward, the acceleration
would have been ay = +g and all our answers would have been
positive.
Solution
ExAmplE 2.7 up-And-down motion in frEE fAll
You throw a ball vertically upward from the roof of a tall building. The ball leaves your hand at a point even with the roof railing
with an upward speed of 15.0 m>s; the ball is then in free fall. On
its way back down, it just misses the railing. Find (a) the ball’s
position and velocity 1.00 s and 4.00 s after leaving your hand;
(b) the ball’s velocity when it is 5.00 m above the railing; (c) the
maximum height reached; (d) the ball’s acceleration when it is at
its maximum height.
Solution
idEntify and SEt up: The words “in free fall” mean that the
acceleration is due to gravity, which is constant. Our target variables are position [in parts (a) and (c)], velocity [in parts (a) and (b)],
and acceleration [in part (d)]. We take the origin at the point
where the ball leaves your hand, and take the positive direction to
be upward (Fig. 2.24). The initial position y0 is zero, the initial
y-velocity v0y is + 15.0 m>s, and the y-acceleration is ay = -g =
- 9.80 m>s2. In part (a), as in Example 2.6, we’ll use Eqs. (2.12)
and (2.8) to find the position and velocity as functions of time.
In part (b) we must find the velocity at a given position (no time
is given), so we’ll use Eq. (2.13).
2.24 Position and velocity of a ball thrown vertically upward.
The ball actually moves straight up and
then straight down; we show
vy = 0
a U-shaped path for clarity.
t = ?
t = 1.00 s, vy = ?
t = ?, vy = ?
t = 0, v0y = 15.0 m>s
t = ?
vy = ?
y
y = ?
y = ?
y = 5.00 m
y = 0
ay = -g
= -9.80 m>s2
t = 4.00 s
vy = ?
y = ?
Continued
52
Chapter 2 Motion along a Straight Line
2.25 (a) Position and (b) velocity as functions of time for a ball
thrown upward with an initial speed of 15.0 m>s.
(a) y-t graph (curvature is
downward because ay = -g
is negative)
y ( m)
15
Before t = 1.53 s the
ball moves upward.
After t = 1.53 s
the ball moves
downward.
10
5
0
1
2
-5
3
4
t (s)
(b) vy-t graph (straight line with
negative slope because ay = -g
is constant and negative)
vy ( m>s)
Before t = 1.53 s
the y-velocity is
positive.
15
10
5
0
-5
1
2
3
-10
-10
-15
-15
-20
-20
-25
t (s)
4
After t = 1.53 s
the y-velocity is
negative.
Figure 2.25 shows the y@t and vy@t graphs for the ball. The y@t
graph is a concave-down parabola that rises and then falls, and the
vy@t graph is a downward-sloping straight line. Note that the ball’s
velocity is zero when it is at its highest point.
(vy is negative) with a speed of 24.2 m>s. Equation (2.13) tells us
that the ball is moving at the initial 15.0@m > s speed as it moves
downward past the launching point and continues to gain speed as
it descends further.
(b) The y-velocity at any position y is given by Eq. (2.13) with
x’s replaced by y’s:
v y2 = v 0y 2 + 2ay 1y - y02 = v 0y 2 + 21- g21y - 02
= 115.0 m>s22 + 21- 9.80 m>s22y
When the ball is 5.00 m above the origin we have y = + 5.00 m, so
v y2 = 115.0 m>s22 + 21-9.80 m>s2215.00 m2 = 127 m2>s2
vy = {11.3 m>s
We get two values of vy because the ball passes through the point
y = +5.00 m twice, once on the way up (so vy is positive) and
once on the way down (so vy is negative) (see Figs. 2.24 and 2.25a).
(c) At the instant at which the ball reaches its maximum height
y1, its y-velocity is momentarily zero: vy = 0. We use Eq. (2.13) to
find y1. With vy = 0, y0 = 0, and ay = -g , we get
0 = v 0y 2 + 2 1-g21y1 - 02
y1 =
ExECutE: (a) The position and y-velocity at time t are given by
Eqs. (2.12) and (2.8) with x’s replaced by y’s:
y = y0 + v0y t + 12 ay t 2 = y0 + v0y t + 12 1-g2t 2
= 102 + 115.0 m>s2t + 12 1- 9.80 m>s22t 2
vy = v0y + ay t = v0y + 1-g2t
= 15.0 m>s + 1- 9.80 m>s22t
When t = 1.00 s, these equations give y = +10.1 m and
vy = + 5.2 m>s. That is, the ball is 10.1 m above the origin (y is
positive) and moving upward (vy is positive) with a speed of
5.2 m>s. This is less than the initial speed because the ball slows
as it ascends. When t = 4.00 s, those equations give y = -18.4 m
and vy = - 24.2 m>s. The ball has passed its highest point and is
18.4 m below the origin (y is negative). It is moving downward
v 0y 2
2g
=
115.0 m>s22
219.80 m>s22
= +11.5 m
(d) Caution A free-fall misconception It’s a common misconception that at the highest point of free-fall motion, where the
velocity is zero, the acceleration is also zero. If this were so, once
the ball reached the highest point it would hang there suspended in
midair! Remember that acceleration is the rate of change of velocity, and the ball’s velocity is continuously changing. At every point,
including the highest point, and at any velocity, including zero, the
acceleration in free fall is always ay = -g = - 9.80 m>s2. ❙
EvaluatE: A useful way to check any free-fall problem is to draw
the y-t and vy@t graphs, as we did in Fig. 2.25. Note that these are
graphs of Eqs. (2.12) and (2.8), respectively. Given the initial position,
initial velocity, and acceleration, you can easily create these graphs
by using a graphing calculator or an online mathematics program.
Solution
ExAmplE 2.8 two solutions or onE?
At what time after being released has the ball in Example 2.7
fallen 5.00 m below the roof railing?
Solution
idEntify and SEt up: We treat this as in Example 2.7, so y0, v0y,
and ay = - g have the same values as there. Now, however, the
target variable is the time at which the ball is at y = -5.00 m.
The best equation to use is Eq. (2.12), which gives the position y as
a function of time t:
y = y0 + v0y t + 12 ay t 2
= y0 + v0y t + 12 1- g2t 2
This is a quadratic equation for t, which we want to solve for the
value of t when y = - 5.00 m.
ExECutE: We rearrange the equation so that it has the standard form
of a quadratic equation for an unknown x, Ax 2 + Bx + C = 0 :
1 12 g 2 t2
+ 1- v0y2t + 1y - y02 = At 2 + Bt + C = 0
By comparison, we identify A = 12 g, B = - v0y, and C = y - y0.
The quadratic formula (see Appendix B) tells us that this equation
has two solutions:
t =
=
=
- B { 2B2 - 4AC
2A
- 1-v0y2 { 21-v0y22 - 4 1 12 g 2 1y - y02
2 1 12 g 2
v0y { 2v 0y2 - 2g 1y - y02
g
2.6 Velocity and position by Integration
Substituting the values y0 = 0, v0y = + 15.0 m>s, g = 9.80 m>s2,
and y = - 5.00 m, we find
t =
115.0 m>s2 { 2115.0 m>s22 - 219.80 m > s221-5.00 m - 02
9.80 m>s2
You can confirm that the numerical answers are t = +3.36 s
and t = - 0.30 s. The answer t = - 0.30 s doesn’t make physical
sense, since it refers to a time before the ball left your hand at
t = 0. So the correct answer is t = + 3.36 s.
evaLuate: Why did we get a second, fictitious solution? The explanation is that constant-acceleration equations like Eq. (2.12) are
based on the assumption that the acceleration is constant for all
values of time, whether positive, negative, or zero. Hence the solution t = - 0.30 s refers to an imaginary moment when a freely
falling ball was 5.00 m below the roof railing and rising to meet
your hand. Since the ball didn’t leave your hand and go into free
fall until t = 0, this result is pure fiction.
53
Repeat these calculations to find the times when the ball is
5.00 m above the origin 1y = + 5.00 m2. The two answers are
t = + 0.38 s and t = +2.68 s. Both are positive values of t, and
both refer to the real motion of the ball after leaving your hand.
At the earlier time the ball passes through y = + 5.00 m moving upward; at the later time it passes through this point moving
downward. [Compare this with part (b) of Example 2.7, and again
refer to Fig. 2.25a.]
You should also solve for the times when y = + 15.0 m. In
this case, both solutions involve the square root of a negative
number, so there are no real solutions. Again Fig. 2.25a shows
why; we found in part (c) of Example 2.7 that the ball’s maximum
height is y = +11.5 m, so it never reaches y = + 15.0 m. While
a quadratic equation such as Eq. (2.12) always has two solutions,
in some situations one or both of the solutions aren’t physically
reasonable.
test your unDerstanDing oF section 2.5 If you toss a ball upward with
a certain initial speed, it falls freely and reaches a maximum height h a time t after it
leaves your hand. (a) If you throw the ball upward with double the initial speed, what
new maximum height does the ball reach? (i) h12 ; (ii) 2h; (iii) 4h; (iv) 8h; (v) 16h.
(b) If you throw the ball upward with double the initial speed, how long does it take to
reach its new maximum height? (i) t>2; (ii) t>12 ; (iii) t; (iv) t12 ; (v) 2t. ❙
2.6 veLocity anD position by integration
This section is intended for students who have already learned a little integral
calculus. In Section 2.4 we analyzed the special case of straight-line motion
with constant acceleration. When ax is not constant, as is frequently the case, the
equations that we derived in that section are no longer valid (Fig. 2.26). But even
when ax varies with time, we can still use the relationship vx = dx>dt to find the
x-velocity vx as a function of time if the position x is a known function of time.
And we can still use ax = dvx >dt to find the x-acceleration ax as a function of
time if the x-velocity vx is a known function of time.
In many situations, however, position and velocity are not known functions of
time, while acceleration is (Fig. 2.27). How can we find the position and velocity
in straight-line motion from the acceleration function ax 1t2?
2.26 When you push a car’s accelerator pedal to the floor-
board, the resulting acceleration is not constant: The greater
the car’s speed, the more slowly it gains additional speed. A
typical car takes twice as long to accelerate from 50 km>h
to 100 km>h as it does to accelerate from 0 to 50 km>h.
2.27 The inertial navigation system (INS) on board a
long-range airliner keeps track of the airliner’s acceleration. Given the airliner’s initial position and velocity before
takeoff, the INS uses the acceleration data to calculate the
airliner’s position and velocity throughout the flight.
54
ChApTer 2 Motion Along a Straight Line
2.28 An ax@t graph for a body whose
x-acceleration is not constant.
ax
Area of this strip = ∆vx
= Change in x-velocity
during time interval ∆t
Figure 2.28 is a graph of x-acceleration versus time for a body whose acceleration is not constant. We can divide the time interval between times t1 and t2 into
many smaller subintervals, calling a typical one ∆t. Let the average x-acceleration
during ∆t be aav@x . From Eq. (2.4) the change in x-velocity ∆vx during ∆t is
∆vx = aav@x ∆t
aav-x
O
t1
∆t
t2
t
Total area under the x-t graph from t1 to t2
= Net change in x-velocity from t1 to t2
Graphically, ∆vx equals the area of the shaded strip with height aav@x and width
∆t—that is, the area under the curve between the left and right sides of ∆t. The
total change in x-velocity from t1 to t2 is the sum of the x-velocity changes ∆vx in
the small subintervals. So the total x-velocity change is represented graphically
by the total area under the ax@t curve between the vertical lines t1 and t2 . (In
Section 2.4 we showed this for the special case in which ax is constant.)
In the limit that all the ∆t>s become very small and they become very large in
number, the value of aav@x for the interval from any time t to t + ∆t approaches
the instantaneous x-acceleration ax at time t. In this limit, the area under the ax@t
curve is the integral of ax (which is in general a function of t) from t1 to t2 . If v1x
is the x-velocity of the body at time t1 and v2x is the velocity at time t2 , then
v2x
v2x - v1x =
Lv1x
t2
dvx =
Lt1
ax dt
(2.15)
The change in the x-velocity vx is the time integral of the x-acceleration ax .
We can carry out exactly the same procedure with the curve of x-velocity versus time. If x1 is a body’s position at time t1 and x2 is its position at time t2 , from
Eq. (2.2) the displacement ∆x during a small time interval ∆t is equal to vav@x ∆t,
where vav@x is the average x-velocity during ∆t. The total displacement x2 - x1
during the interval t2 - t1 is given by
x2
x2 - x1 =
Lx1
t2
dx =
Lt1
vx dt
(2.16)
The change in position x—that is, the displacement—is the time integral of
x-velocity vx. Graphically, the displacement between times t1 and t2 is the area
under the vx@t curve between those two times. [This is the same result that we
obtained in Section 2.4 for the special case in which vx is given by Eq. (2.8).]
If t1 = 0 and t2 is any later time t, and if x0 and v0x are the position and velocity, respectively, at time t = 0, then we can rewrite Eqs. (2.15) and (2.16) as
x-velocity of a
particle at time t
x-velocity of the particle at time 0
vx = v0x +
L0
t
ax dt
(2.17)
Integral of the x-acceleration of the particle from time 0 to time t
Position of a
particle at time t
Position of the particle at time 0
x = x0 +
L0
t
vx dt
(2.18)
Integral of the x-velocity of the particle from time 0 to time t
If we know the x-acceleration ax as a function of time and we know the initial
velocity v0x , we can use Eq. (2.17) to find the x-velocity vx at any time; in other
words, we can find vx as a function of time. Once we know this function, and
given the initial position x0 , we can use Eq. (2.18) to find the position x at any
time.
2.6 Velocity and position by Integration
55
Solution
ExamplE 2.9 moTion wiTh changing accElEraTion
Sally is driving along a straight highway in her 1965 Mustang. At
t = 0, when she is moving at 10 m>s in the positive x-direction,
she passes a signpost at x = 50 m. Her x-acceleration as a function of time is
ax = 2.0 m>s2 - 10.10 m>s32t
(a) Find her x-velocity vx and position x as functions of time.
(b) When is her x-velocity greatest? (c) What is that maximum
x-velocity? (d) Where is the car when it reaches that maximum
x-velocity?
soLution
iDentiFy and set up: The x-acceleration is a function of time, so
we cannot use the constant-acceleration formulas of Section 2.4.
Instead, we use Eq. (2.17) to obtain an expression for vx as a
function of time, and then use that result in Eq. (2.18) to find an
expression for x as a function of t. We’ll then be able to answer a
variety of questions about the motion.
Figure 2.29 shows graphs of ax, vx, and x as functions of time as
given by the previous equations. Note that for any time t, the slope
of the vx@t graph equals the value of ax and the slope of the x-t
graph equals the value of vx.
(b) The maximum value of vx occurs when the x-velocity stops
increasing and begins to decrease. At that instant, dvx >dt = ax = 0.
So we set the expression for ax equal to zero and solve for t:
0 = 2.0 m>s2 - 10.10 m>s32t
t =
vmax@x = 10 m>s + 12.0 m>s22120 s2 - 12 10.10 m>s32120 s22
= 30 m>s
(d) To find the car’s position at the time that we found in part (b),
we substitute t = 20 s into the expression for x from part (a):
x = 50 m + 110 m>s2120 s2 + 12 12.0 m>s22120 s22
x-velocity is v0x = 10 m>s. To use Eq. (2.17), we note that the integral of t n (except for n = - 1) is 1 t n dt = n +1 1 t n + 1. Hence
vx = 10 m>s +
L0
= 517 m
hand graph shows that ax is positive between t = 0 and t = 20 s
and negative after that. It is zero at t = 20 s, the time at which vx
is maximum (the high point in the middle graph). The car speeds
up until t = 20 s (because vx and ax have the same sign) and slows
down after t = 20 s (because vx and ax have opposite signs).
Since vx is maximum at t = 20 s, the x-t graph (the righthand graph in Fig. 2.29) has its maximum positive slope at this
time. Note that the x-t graph is concave up (curved upward) from
t = 0 to t = 20 s, when ax is positive. The graph is concave down
(curved downward) after t = 20 s, when ax is negative.
Now we use Eq. (2.18) to find x as a function of t:
L0
- 16 10.10 m>s32120 s23
evaLuate: Figure 2.29 helps us interpret our results. The left-
32.0 m>s2 - 10.10 m>s32t4 dt
= 10 m>s + 12.0 m>s22t - 12 10.10 m>s32t 2
x = 50 m +
= 20 s
0.10 m>s3
(c) We find the maximum x-velocity by substituting t = 20 s,
the time from part (b) when velocity is maximum, into the equation for vx from part (a):
execute: (a) At t = 0, Sally’s position is x0 = 50 m and her
t
2.0 m>s2
t
310 m>s + 12.0 m>s22t - 12 10.10 m>s32t 24 dt
= 50 m + 110 m>s2t + 12 12.0 m>s22t 2 - 16 10.10 m>s32t 3
2.29 The position, velocity, and acceleration of the car in Example 2.9 as functions of time.
Can you show that if this motion continues, the car will stop at t = 44.5 s?
x (m)
800
vx (m>s)
30
ax (m>s2)
x-acceleration is
positive before t = 20 s.
2.0
20
1.0
O
-1.0
10
5
10 15 20 25
x-acceleration is
negative after t = 20 s.
30
t (s)
O
x-t graph curves
upward before
t = 20 s.
600
x-velocity
increases before
t = 20 s.
5
10
15
x-velocity
decreases after
t = 20 s.
20
25
30
t (s)
test your unDerstanDing oF section 2.6 If the x-acceleration ax of an
object moving in straight-line motion is increasing with time, will the vx@t graph be (i) a
straight line, (ii) concave up (i.e., with an upward curvature), or (iii) concave down
(i.e., with a downward curvature)? ❙
x-t graph curves
downward after
t = 20 s.
400
200
O
5
10
15
20
25
30
t (s)
Chapter 2 Motion along a Straight Line
2 Summary
x-acceleration aav@x during a time interval ∆t is equal
to the change in velocity ∆vx = v2x - v1x during
that time interval divided by ∆t. The instantaneous
x-acceleration ax is the limit of aav@x as ∆t goes to
zero, or the derivative of vx with respect to t. (See
Examples 2.2 and 2.3.)
(2.3)
p1
x1
t1
O
aav@x =
∆vx
v2x - v1x
=
∆t
t2 - t1
(2.4)
∆vx
dvx
=
∆t
dt
(2.5)
ax = lim
∆t S 0
When the x-acceleration is constant, four equations
relate the position x and the x-velocity vx at any time t
to the initial position x0 , the initial x-velocity v0x (both
measured at time t = 0), and the x-acceleration ax . (See
Examples 2.4 and 2.5.)
p2
-x
p1
t = ∆t
v x2 = v 0x 2 + 2ax 1x - x02
(2.13)
t = 2∆t
x - x0 =
+ vx2t
(2.14)
x
v
(2.12)
a
x
0
t = 3∆t
t = 4∆t
t
t2
a
0
x = x0 + v0x t + 12 ax t 2
a
e = x
∆t = t2 - t1
v
t = 0
a av
Slop
t1
Constant x-acceleration only:
1
2 1v0x
=
pe
o
Sl
(2.8)
t
t2
∆t = t2 - t1
v2x
v1x
vx = v0x + ax t
= vx
pe
Slo
vx
O
Straight-line motion with constant acceleration:
p2
∆x = x2 - x1
∆t S 0
x
x2
av
-x
∆x
dx
=
∆t
dt
vx = lim
(2.2)
∆vx = v2x - v1x
Average and instantaneous x-acceleration: The average
x2 - x1
∆x
=
∆t
t2 - t1
v
we describe its position with respect to an origin O by
means of a coordinate such as x. The particle’s average
x-velocity vav@x during a time interval ∆t = t2 - t1 is
equal to its displacement ∆x = x2 - x1 divided by ∆t.
The instantaneous x-velocity vx at any time t is equal to
the average x-velocity over the time interval from t to
t + ∆t in the limit that ∆t goes to zero. Equivalently,
vx is the derivative of the position function with respect
to time. (See Example 2.1.)
vav@x =
=
Straight-line motion, average and instantaneous
x-velocity: When a particle moves along a straight line,
SolutionS to all exampleS
pe
Chapter
Sl
o
56
v
a
x
0
v
a
x
0
v
0
a
x
Freely falling bodies: Free fall is a case of motion with
constant acceleration. The magnitude of the acceleration due to gravity is a positive quantity, g. The
acceleration of a body in free fall is always downward.
(See Examples 2.6–2.8.)
Straight-line motion with varying acceleration:
When the acceleration is not constant but is a known
function of time, we can find the velocity and position
as functions of time by integrating the acceleration
function. (See Example 2.9.)
ay = -g
= -9.80 m>s2
vx = v0x +
x = x0 +
L0
L0
t
ax dt
ax
(2.17)
t
vx dt
(2.18)
aav-x
O
t1
∆t
t2
t
Discussion Questions
Solution
BriDging proBlEm
57
ThE Fall oF a SupErhEro
The superhero Green Lantern steps from the top of a tall building.
He falls freely from rest to the ground, falling half the total distance
to the ground during the last 1.00 s of his fall (Fig. 2.30). What is
the height h of the building?
2.30 Our sketch for this problem.
Solution GuidE
idEntify and SEt up
1. You’re told that Green Lantern falls freely from rest. What does
this imply about his acceleration? About his initial velocity?
2. Choose the direction of the positive y-axis. It’s easiest to make
the same choice we used for freely falling objects in Section 2.5.
3. You can divide Green Lantern’s fall into two parts: from the
top of the building to the halfway point and from the halfway
point to the ground. You know that the second part of the fall
lasts 1.00 s. Decide what you would need to know about Green
Lantern’s motion at the halfway point in order to solve for the
target variable h. Then choose two equations, one for the first
part of the fall and one for the second part, that you’ll use together to find an expression for h. (There are several pairs of
equations that you could choose.)
ExECutE
4. Use your two equations to solve for the height h. Heights are
always positive numbers, so your answer should be positive.
h = ?
Falls last h>2
in 1.00 s
EValuatE
5. To check your answer for h, use one of the free-fall equations
to find how long it takes Green Lantern to fall (i) from the top
of the building to half the height and (ii) from the top of the
building to the ground. If your answer for h is correct, time
(ii) should be 1.00 s greater than time (i). If it isn’t, go back and
look for errors in how you found h.
Problems
For assigned homework and other learning materials, go to MasteringPhysics®.
., .., ...: Difficulty levels. CP: Cumulative problems incorporating material from earlier chapters. CALC: Problems requiring calculus.
DATA: Problems involving real data, scientific evidence, experimental design, and/or statistical reasoning. BIO: Biosciences problems.
diSCuSSion QuEStionS
Q2.1 Does the speedometer of a car measure speed or velocity?
Explain.
Q2.2 The black dots at the top of Fig. Q2.2 represent a series of
high-speed photographs of an insect flying in a straight line from
left to right (in the positive x-direction). Which of the graphs in
Fig. Q2.2 most plausibly depicts this insect’s motion?
Figure Q2.2
vx
ax
t
O
(a)
t
O
(b)
vx
vx
x
t
O
(c)
t
O
(d)
t
O
(e)
Q2.3 Can an object with constant acceleration reverse its direction of travel? Can it reverse its direction twice? In both cases,
explain your reasoning.
Q2.4 Under what conditions is average velocity equal to instantaneous velocity?
Q2.5 Is it possible for an object to be (a) slowing down while its
acceleration is increasing in magnitude; (b) speeding up while its
acceleration is decreasing? In both cases, explain your reasoning.
Q2.6 Under what conditions does the magnitude of the average
velocity equal the average speed?
Q2.7 When a Dodge Viper is at Elwood’s Car Wash, a BMW
Z3 is at Elm and Main. Later, when the Dodge reaches Elm and
Main, the BMW reaches Elwood’s Car Wash. How are the cars’
average velocities between these two times related?
Q2.8 A driver in Massachusetts was sent to traffic court for
speeding. The evidence against the driver was that a policewoman
observed the driver’s car alongside a second car at a certain moment, and the policewoman had already clocked the second car
going faster than the speed limit. The driver argued, “The second
car was passing me. I was not speeding.” The judge ruled against
the driver because, in the judge’s words, “If two cars were side by
side, both of you were speeding.” If you were a lawyer representing the accused driver, how would you argue this case?
Q2.9 Can you have zero displacement and nonzero average velocity? Zero displacement and nonzero velocity? Illustrate your
answers on an x-t graph.
Q2.10 Can you have zero acceleration and nonzero velocity? Use
a vx@t graph to explain.
Q2.11 Can you have zero velocity and nonzero average acceleration? Zero velocity and nonzero acceleration? Use a vx@t graph to
explain, and give an example of such motion.
Q2.12 An automobile is traveling west. Can it have a velocity
toward the west and at the same time have an acceleration toward
the east? Under what circumstances?
58
Chapter 2 Motion along a Straight Line
Q2.13 The official’s truck in Fig. 2.2 is at x1 = 277 m at t1 =
16.0 s and is at x2 = 19 m at t2 = 25.0 s. (a) Sketch two different
possible x-t graphs for the motion of the truck. (b) Does the average
velocity vav@x during the time interval from t1 to t2 have the same
value for both of your graphs? Why or why not?
Q2.14 Under constant acceleration the average velocity of a particle is half the sum of its initial and final velocities. Is this still
true if the acceleration is not constant? Explain.
Q2.15 You throw a baseball straight up in the air so that it rises
to a maximum height much greater than your height. Is the magnitude of the ball’s acceleration greater while it is being thrown or
after it leaves your hand? Explain.
Q2.16 Prove these statements: (a) As long as you can ignore the
effects of the air, if you throw anything vertically upward, it will
have the same speed when it returns to the release point as when it
was released. (b) The time of flight will be twice the time it takes
to get to its highest point.
Q2.17 A dripping water faucet steadily releases drops 1.0 s apart.
As these drops fall, does the distance between them increase, decrease, or remain the same? Prove your answer.
Q2.18 If you know the initial position and initial velocity of a vehicle and have a record of the acceleration at each instant, can you
compute the vehicle’s position after a certain time? If so, explain
how this might be done.
Q2.19 From the top of a tall building, you throw one ball straight
up with speed v0 and one ball straight down with speed v0.
(a) Which ball has the greater speed when it reaches the ground?
(b) Which ball gets to the ground first? (c) Which ball has a
greater displacement when it reaches the ground? (d) Which ball
has traveled the greater distance when it hits the ground?
Q2.20 You run due east at a constant speed of 3.00 m>s for a
distance of 120.0 m and then continue running east at a constant
speed of 5.00 m>s for another 120.0 m. For the total 240.0-m run,
is your average velocity 4.00 m>s, greater than 4.00 m>s, or less
than 4.00 m>s? Explain.
Q2.21 An object is thrown straight up into the air and feels no
air resistance. How can the object have an acceleration when it has
stopped moving at its highest point?
Q2.22 When you drop an object from a certain height, it takes
time T to reach the ground with no air resistance. If you dropped
it from three times that height, how long (in terms of T) would it
take to reach the ground?
exercises
Section 2.1 Displacement, Time,
and average Velocity
2.1 . A car travels in the + x-direction on a straight and level
road. For the first 4.00 s of its motion, the average velocity of the
car is vav@x = 6.25 m>s. How far does the car travel in 4.00 s?
2.2 .. In an experiment, a shearwater (a seabird) was taken from
its nest, flown 5150 km away, and released. The bird found its way
back to its nest 13.5 days after release. If we place the origin at the
nest and extend the + x@axis to the release point, what was the
bird’s average velocity in m>s (a) for the return flight and (b) for
the whole episode, from leaving the nest to returning?
2.3 .. Trip Home. You normally drive on the freeway between San Diego and Los Angeles at an average speed of
105 km>h 165 mi>h2, and the trip takes 1 h and 50 min. On a
Friday afternoon, however, heavy traffic slows you down and you
drive the same distance at an average speed of only 70 km>h
143 mi>h2. How much longer does the trip take?
.. From Pillar to Post. Starting from a pillar, you run
200 m east (the +x@direction) at an average speed of 5.0 m>s and
then run 280 m west at an average speed of 4.0 m>s to a post.
Calculate (a) your average speed from pillar to post and (b) your
average velocity from pillar to post.
2.5 . Starting from the front door of a ranch house, you walk
60.0 m due east to a windmill, turn around, and then slowly walk
40.0 m west to a bench, where you sit and watch the sunrise. It
takes you 28.0 s to walk from the house to the windmill and then
36.0 s to walk from the windmill to the bench. For the entire trip
from the front door to the bench, what are your (a) average velocity
and (b) average speed?
2.6 .. A Honda Civic travels in a straight line along a road. The
car’s distance x from a stop sign is given as a function of time t
by the equation x1t2 = at 2 - bt 3, where a = 1.50 m>s2 and
b = 0.0500 m>s3. Calculate the average velocity of the car for
each time interval: (a) t = 0 to t = 2.00 s; (b) t = 0 to t = 4.00 s;
(c) t = 2.00 s to t = 4.00 s.
2.4
Section 2.2 Instantaneous Velocity
. CALC A car is stopped at a traffic light. It then travels
along a straight road such that its distance from the light is given
by x1t2 = bt 2 - ct 3, where b = 2.40 m>s2 and c = 0.120 m>s3.
(a) Calculate the average velocity of the car for the time interval
t = 0 to t = 10.0 s. (b) Calculate the instantaneous velocity of the
car at t = 0, t = 5.0 s, and t = 10.0 s. (c) How long after starting
from rest is the car again at rest?
2.8 . CALC A bird is flying due east. Its distance from a tall building is given by x1t2 = 28.0 m + 112.4 m>s2t - 10.0450 m>s32t 3.
What is the instantaneous velocity of the bird when t = 8.00 s?
2.9 .. A ball moves in a
straight line (the x-axis). The Figure e2.9
graph in Fig. E2.9 shows this
vx (m/s)
ball’s velocity as a function
3.0
of time. (a) What are the ball’s
average speed and average
2.0
velocity during the first 3.0 s?
1.0
(b) Suppose that the ball moved
in such a way that the graph segt (s)
ment after 2.0 s was -3.0 m>s
1.0
2.0
3.0
O
instead of +3.0 m>s. Find the
ball’s average speed and average
velocity in this case.
2.10 .. A physics professor leaves her house and walks along the
sidewalk toward campus. After 5 min it starts to rain, and she returns home. Her distance from her house as a function of time is
shown in Fig. E2.10. At which of the labeled points is her velocity
(a) zero? (b) constant and positive? (c) constant and negative?
(d) increasing in magnitude? (e) decreasing in magnitude?
2.7
Figure e2.10
x (m)
IV
400
III
300
V
200
II
100
I
O
1
2
3
4
5
6
7
8
t (min)
exercises
2.11 .. A test car travels in a
straight line along the x-axis.
The graph in Fig. E2.11
shows the car’s position x as a
function of time. Find its instantaneous velocity at points
A through G.
Figure e2.11
x ( m)
C
B
40
30 A
20
10
O
D
E
1 2 3 4 5 6 7 8 9 10
F G
t (s)
Section 2.3 average and Instantaneous acceleration
2.12
.
Figure E2.12 shows the velocity of a solar-powered car
as a function of time. The driver accelerates from a stop sign,
cruises for 20 s at a constant speed of 60 km / h, and then brakes to
come to a stop 40 s after leaving the stop sign. (a) Compute
the average acceleration during these time intervals: (i) t = 0
to t = 10 s; (ii) t = 30 s to t = 40 s; (iii) t = 10 s to t = 30 s;
(iv) t = 0 to t = 40 s. (b) What is the instantaneous acceleration
at t = 20 s and at t = 35 s ?
Figure e2.12
vx (km>h)
60
50
40
30
59
starting does it take the turtle to return to its starting point? (d) At
what times t is the turtle a distance of 10.0 cm from its starting
point? What is the velocity (magnitude and direction) of the
turtle at each of those times? (e) Sketch graphs of x versus t,
vx versus t, and ax versus t, for the time interval t = 0 to t = 40 s.
2.16 . An astronaut has left the International Space Station to
test a new space scooter. Her partner measures the following velocity changes, each taking place in a 10-s interval. What are the
magnitude, the algebraic sign, and the direction of the average acceleration in each interval? Assume that the positive direction is to
the right. (a) At the beginning of the interval, the astronaut is moving toward the right along the x-axis at 15.0 m>s, and at the end of
the interval she is moving toward the right at 5.0 m>s. (b) At the
beginning she is moving toward the left at 5.0 m>s, and at the end
she is moving toward the left at 15.0 m>s. (c) At the beginning she
is moving toward the right at 15.0 m>s, and at the end she is moving toward the left at 15.0 m>s.
2.17 . CALC A car’s velocity as a function of time is given by
vx1t2 = a + bt 2, where a = 3.00 m>s and b = 0.100 m>s3.
(a) Calculate the average acceleration for the time interval t = 0
to t = 5.00 s. (b) Calculate the instantaneous acceleration for
t = 0 and t = 5.00 s. (c) Draw vx@t and ax@t graphs for the car’s
motion between t = 0 and t = 5.00 s.
2.18 .. CALC The position of the front bumper of a test car
under microprocessor control is given by x 1t2 = 2.17 m +
14.80 m>s22t 2 - 10.100 m>s62t 6. (a) Find its position and acceleration at the instants when the car has zero velocity. (b) Draw x-t,
vx@t, and ax@t graphs for the motion of the bumper between t = 0
and t = 2.00 s.
20
Section 2.4 Motion with Constant acceleration
10
O
5 10 15 20 25 30 35 40
t (s)
. The Fastest (and Most Expensive) Car! The table shows
test data for the Bugatti Veyron Super Sport, the fastest street car
made. The car is moving in a straight line (the x-axis).
2.13
Time (s)
0
2.1
20.0
53
Speed (mi , h)
0
60
200
253
(a) Sketch a vx@t graph of this car’s velocity (in mi>h) as a function
of time. Is its acceleration constant? (b) Calculate the car’s average
acceleration (in m>s2) between (i) 0 and 2.1 s; (ii) 2.1 s and 20.0 s;
(iii) 20.0 s and 53 s. Are these results consistent with your graph in
part (a)? (Before you decide to buy this car, it might be helpful to
know that only 300 will be built, it runs out of gas in 12 minutes at
top speed, and it costs more than $1.5 million!)
2.14 .. CALC A race car starts from rest and travels east along
a straight and level track. For the first 5.0 s of the car’s motion,
the eastward component of the car’s velocity is given by
vx1t2 = 10.860 m>s32t 2. What is the acceleration of the car when
vx = 12.0 m>s?
2.15 . CALC A turtle crawls along a straight line, which we will
call the x-axis with the positive direction to the right. The equation for the turtle’s position as a function of time is x 1t2 =
50.0 cm + 12.00 cm>s2t - 10.0625 cm>s22t 2. (a) Find the turtle’s
initial velocity, initial position, and initial acceleration. (b) At
what time t is the velocity of the turtle zero? (c) How long after
2.19 .. An antelope moving with constant acceleration covers
the distance between two points 70.0 m apart in 6.00 s. Its speed
as it passes the second point is 15.0 m>s. What are (a) its speed at
the first point and (b) its acceleration?
2.20 .. BIO Blackout? A jet fighter pilot wishes to accelerate
from rest at a constant acceleration of 5g to reach Mach 3 (three
times the speed of sound) as quickly as possible. Experimental
tests reveal that he will black out if this acceleration lasts for more
than 5.0 s. Use 331 m>s for the speed of sound. (a) Will the period of
acceleration last long enough to cause him to black out? (b) What
is the greatest speed he can reach with an acceleration of 5g before
he blacks out?
2.21 . A Fast Pitch. The fastest measured pitched baseball
left the pitcher’s hand at a speed of 45.0 m>s. If the pitcher was in
contact with the ball over a distance of 1.50 m and produced constant acceleration, (a) what acceleration did he give the ball, and
(b) how much time did it take him to pitch it?
2.22 .. A Tennis Serve. In the fastest measured tennis serve,
the ball left the racquet at 73.14 m>s. A served tennis ball is typically in contact with the racquet for 30.0 ms and starts from rest.
Assume constant acceleration. (a) What was the ball’s acceleration
during this serve? (b) How far did the ball travel during the serve?
2.23 .. BIO Automobile Air Bags. The human body can survive an acceleration trauma incident (sudden stop) if the magnitude of the acceleration is less than 250 m>s2. If you are in an
automobile accident with an initial speed of 105 km>h 165 mi>h2
and are stopped by an airbag that inflates from the dashboard,
over what distance must the airbag stop you for you to survive
the crash?
60
Chapter 2 Motion along a Straight Line
. BIO A pilot who accelerates at more than 4g begins to
“gray out” but doesn’t completely lose consciousness. (a) Assuming
constant acceleration, what is the shortest time that a jet pilot
starting from rest can take to reach Mach 4 (four times the speed
of sound) without graying out? (b) How far would the plane travel
during this period of acceleration? (Use 331 m>s for the speed of
sound in cold air.)
2.25 . BIO Air-Bag Injuries. During an auto accident, the
vehicle’s air bags deploy and slow down the passengers more gently than if they had hit the windshield or steering wheel. According
to safety standards, air bags produce a maximum acceleration of
60g that lasts for only 36 ms (or less). How far (in meters) does a
person travel in coming to a complete stop in 36 ms at a constant
acceleration of 60g?
2.26 . BIO Prevention of Hip Fractures. Falls resulting in hip
fractures are a major cause of injury and even death to the elderly.
Typically, the hip’s speed at impact is about 2.0 m>s. If this can be
reduced to 1.3 m>s or less, the hip will usually not fracture. One
way to do this is by wearing elastic hip pads. (a) If a typical pad is
5.0 cm thick and compresses by 2.0 cm during the impact of a fall,
what constant acceleration (in m>s2 and in g’s) does the hip undergo
to reduce its speed from 2.0 m>s to 1.3 m>s? (b) The acceleration
you found in part (a) may seem rather large, but to assess its effects
on the hip, calculate how long it lasts.
2.27 . BIO Are We Martians? It has been suggested, and not
facetiously, that life might have originated on Mars and been carried to the earth when a meteor hit Mars and blasted pieces of
rock (perhaps containing primitive life) free of the Martian surface. Astronomers know that many Martian rocks have come to
the earth this way. (For instance, search the Internet for “ALH
84001.”) One objection to this idea is that microbes would have
had to undergo an enormous lethal acceleration during the impact.
Let us investigate how large such an acceleration might be. To
escape Mars, rock fragments would have to reach its escape
velocity of 5.0 km>s, and that would most likely happen over
a distance of about 4.0 m during the meteor impact. (a) What
would be the acceleration (in m>s2 and g’s) of such a rock fragment, if the acceleration is constant? (b) How long would this
acceleration last? (c) In tests, scientists have found that over
40% of Bacillus subtilis bacteria survived after an acceleration
of 450,000g. In light of your answer to part (a), can we rule out
the hypothesis that life might have been blasted from Mars to
the earth?
2.28 . Entering the Freeway. A car sits on an entrance ramp
to a freeway, waiting for a break in the traffic. Then the driver accelerates with constant acceleration along the ramp and onto the
freeway. The car starts from rest, moves in a straight line, and has
a speed of 20 m>s 145 mi>h2 when it reaches the end of the 120-mlong ramp. (a) What is the acceleration of the car? (b) How much
time does it take the car to travel the length of the ramp? (c) The
traffic on the freeway is moving at a constant speed of 20 m>s.
What distance does the traffic travel while the car is moving the
length of the ramp?
2.29 .. At launch a rocket ship weighs 4.5 million pounds.
When it is launched from rest, it takes 8.00 s to reach 161 km>h; at
the end of the first 1.00 min, its speed is 1610 km>h. (a) What is
the average acceleration (in m>s2) of the rocket (i) during the first
8.00 s and (ii) between 8.00 s and the end of the first 1.00 min?
(b) Assuming the acceleration is constant during each time interval (but not necessarily the same in both intervals), what distance
does the rocket travel (i) during the first 8.00 s and (ii) during the
interval from 8.00 s to 1.00 min?
2.24
2.30 .. A cat walks in a straight line, which we shall call the
x-axis, with the positive direction to the right. As an observant physicist, you make measurements of this cat’s motion and construct
a graph of the feline’s velocity as a function of time (Fig. E2.30).
(a) Find the cat’s velocity at t = 4.0 s and at t = 7.0 s.
(b) What is the cat’s acceleration at t = 3.0 s? At t = 6.0 s ? At
t = 7.0 s ? (c) What distance does the cat move during the first
4.5 s? From t = 0 to t = 7.5 s ? (d) Assuming that the cat started
at the origin, sketch clear graphs of the cat’s acceleration and position as functions of time.
Figure e2.30
vx (cm>s)
8
6
4
2
O
2.31
1
2
3
4
5
6
7
t (s)
..
The graph in Fig. E2.31 shows the velocity of a motorcycle police officer plotted as a function of time. (a) Find the
instantaneous acceleration at t = 3 s, t = 7 s, and t = 11 s.
(b) How far does the officer go in the first 5 s? The first 9 s? The
first 13 s?
Figure e2.31
vx (m>s)
50
45
40
35
30
25
20
15
10
5
O
2 4 6 8 10 12 14
t (s)
2.32 . Two cars, A and B, move Figure e2.32
along the x-axis. Figure E2.32 is
x (m)
a graph of the positions of A and
25
B versus time. (a) In motion diaA
20
grams (like Figs. 2.13b and
15
B
2.14b), show the position, veloc10
ity, and acceleration of each of
5
the two cars at t = 0, t = 1 s,
and t = 3 s. (b) At what time(s),
t (s)
O
1
2
3
4
if any, do A and B have the same
position? (c) Graph velocity versus time for both A and B. (d) At
what time(s), if any, do A and B have the same velocity? (e) At
what time(s), if any, does car A pass car B? (f) At what time(s), if
any, does car B pass car A?
2.33 .. A small block has constant acceleration as it slides down
a frictionless incline. The block is released from rest at the top of
the incline, and its speed after it has traveled 6.80 m to the bottom
of the incline is 3.80 m>s. What is the speed of the block when it is
3.40 m from the top of the incline?
exercises
. At the instant the traffic light turns green, a car that has
been waiting at an intersection starts ahead with a constant acceleration of 2.80 m>s2. At the same instant a truck, traveling with a
constant speed of 20.0 m>s, overtakes and passes the car. (a) How
far beyond its starting point does the car overtake the truck?
(b) How fast is the car traveling when it overtakes the truck?
(c) Sketch an x-t graph of the motion of both vehicles. Take x = 0
at the intersection. (d) Sketch a vx@t graph of the motion of both
vehicles.
2.34
Section 2.5 Freely Falling bodies
2.35
..
(a) If a flea can jump straight up to a height of 0.440 m,
what is its initial speed as it leaves the ground? (b) How long is it
in the air?
2.36 .. A small rock is thrown vertically upward with a speed of
22.0 m>s from the edge of the roof of a 30.0-m-tall building. The
rock doesn’t hit the building on its way back down and lands on
the street below. Ignore air resistance. (a) What is the speed of the
rock just before it hits the street? (b) How much time elapses from
when the rock is thrown until it hits the street?
2.37 . A juggler throws a bowling pin straight up with an initial
speed of 8.20 m>s. How much time elapses until the bowling pin
returns to the juggler’s hand?
2.38 .. You throw a glob of putty straight up toward the ceiling,
which is 3.60 m above the point where the putty leaves your hand.
The initial speed of the putty as it leaves your hand is 9.50 m>s.
(a) What is the speed of the putty just before it strikes the ceiling?
(b) How much time from when it leaves your hand does it take the
putty to reach the ceiling?
2.39 .. A tennis ball on Mars, where the acceleration due to
gravity is 0.379g and air resistance is negligible, is hit directly upward and returns to the same level 8.5 s later. (a) How high above
its original point did the ball go? (b) How fast was it moving just
after it was hit? (c) Sketch graphs of the ball’s vertical position,
vertical velocity, and vertical acceleration as functions of time
while it’s in the Martian air.
2.40 .. Touchdown on the Moon. A lunar lander is making
its descent to Moon Base I (Fig. E2.40). The lander descends
slowly under the retro-thrust of its descent engine. The engine is
cut off when the lander is 5.0 m above the surface and has a downward speed of 0.8 m>s.With the engine off, the lander is in free
fall. What is the speed of the lander just before it touches the surface? The acceleration due to gravity on the moon is 1.6 m>s2.
Figure e2.40
5.0 m
2.41 .. A Simple Reaction-Time Test. A meter stick is held
vertically above your hand, with the lower end between your
thumb and first finger. When you see the meter stick released, you
grab it with those two fingers. You can calculate your reaction
time from the distance the meter stick falls, read directly from the
point where your fingers grabbed it. (a) Derive a relationship for
61
your reaction time in terms of this measured distance, d. (b) If the
measured distance is 17.6 cm, what is your reaction time?
2.42 .. A brick is dropped (zero initial speed) from the roof of a
building. The brick strikes the ground in 1.90 s. You may ignore
air resistance, so the brick is in free fall. (a) How tall, in meters, is
the building? (b) What is the magnitude of the brick’s velocity just
before it reaches the ground? (c) Sketch ay@t, vy@t, and y-t graphs
for the motion of the brick.
2.43 .. Launch Failure. A 7500-kg rocket blasts off vertically from the launch pad with a constant upward acceleration of
2.25 m>s2 and feels no appreciable air resistance. When it has
reached a height of 525 m, its engines suddenly fail; the only force
acting on it is now gravity. (a) What is the maximum height this
rocket will reach above the launch pad? (b) How much time will
elapse after engine failure before the rocket comes crashing down
to the launch pad, and how fast will it be moving just before it
crashes? (c) Sketch ay@t, vy@t, and y-t graphs of the rocket’s motion
from the instant of blast-off to the instant just before it strikes the
launch pad.
2.44 .. A hot-air balloonist, risFigure e2.44
ing vertically with a constant vev = 5.00 m>s
locity of magnitude 5.00 m>s,
releases a sandbag at an instant
when the balloon is 40.0 m above
the ground (Fig. E2.44). After the
sandbag is released, it is in free
fall. (a) Compute the position and
velocity of the sandbag at 0.250 s
and 1.00 s after its release. (b) How
many seconds after its release does
the bag strike the ground? (c) With
what magnitude of velocity does
it strike the ground? (d) What is
the greatest height above the
ground that the sandbag reaches?
40.0 m to ground
(e) Sketch ay@t, vy@t, and y-t graphs
for the motion.
2.45 . BIO The rocket-driven sled Sonic Wind No. 2, used for
investigating the physiological effects of large accelerations, runs
on a straight, level track 1070 m (3500 ft) long. Starting from rest,
it can reach a speed of 224 m>s1500 mi>h2 in 0.900 s. (a) Compute
the acceleration in m>s2, assuming that it is constant. (b) What is
the ratio of this acceleration to that of a freely falling body (g)?
(c) What distance is covered in 0.900 s? (d) A magazine article
states that at the end of a certain run, the speed of the sled decreased from 283 m>s 1632 mi>h2 to zero in 1.40 s and that during
this time the magnitude of the acceleration was greater than 40g.
Are these figures consistent?
2.46 . An egg is thrown nearly vertically upward from a point
near the cornice of a tall building. The egg just misses the cornice
on the way down and passes a point 30.0 m below its starting
point 5.00 s after it leaves the thrower’s hand. Ignore air resistance.
(a) What is the initial speed of the egg? (b) How high does it rise
above its starting point? (c) What is the magnitude of its velocity
at the highest point? (d) What are the magnitude and direction of
its acceleration at the highest point? (e) Sketch ay@t, vy@t, and y-t
graphs for the motion of the egg.
2.47 .. A 15-kg rock is dropped from rest on the earth and
reaches the ground in 1.75 s. When it is dropped from the same
height on Saturn’s satellite Enceladus, the rock reaches the
ground in 18.6 s. What is the acceleration due to gravity on
Enceladus?
62
ChApTer 2 Motion Along a Straight Line
. A large boulder is ejected vertically upward from a volcano with an initial speed of 40.0 m>s. Ignore air resistance. (a) At
what time after being ejected is the boulder moving at 20.0 m>s
upward? (b) At what time is it moving at 20.0 m>s downward?
(c) When is the displacement of the boulder from its initial position
zero? (d) When is the velocity of the boulder zero? (e) What are
the magnitude and direction of the acceleration while the boulder
is (i) moving upward? (ii) Moving downward? (iii) At the highest
point? (f) Sketch ay@t, vy@t, and y-t graphs for the motion.
2.49 .. You throw a small rock straight up from the edge of a
highway bridge that crosses a river. The rock passes you on its
way down, 6.00 s after it was thrown. What is the speed of the
rock just before it reaches the water 28.0 m below the point where
the rock left your hand? Ignore air resistance.
2.50 .. CALC A small object moves along the x-axis with acceleration ax1t2 = - 10.0320 m>s32115.0 s - t2. At t = 0 the object
is at x = - 14.0 m and has velocity v0x = 8.00 m>s. What is the
x-coordinate of the object when t = 10.0 s?
2.48
Section 2.6 Velocity and Position by Integration
. CALC A rocket starts from rest and moves upward from
the surface of the earth. For the first 10.0 s of its motion, the vertical acceleration of the rocket is given by ay = 12.80 m>s32t, where
the + y-direction is upward. (a) What is the height of the rocket
above the surface of the earth at t = 10.0 s? (b) What is the speed
of the rocket when it is 325 m above the surface of the earth?
2.52 .. CALC The acceleration of a bus is given by ax1t2 = at,
where a = 1.2 m>s3. (a) If the bus’s velocity at time t = 1.0 s
is 5.0 m>s, what is its velocity at time t = 2.0 s ? (b) If the
bus’s position at time t = 1.0 s is 6.0 m, what is its position at
time t = 2.0 s ? (c) Sketch ay@t, vy@t, and x-t graphs for the
motion.
2.53 .. CALC The acceleration of a motorcycle is given by
ax1t2 = At - Bt 2, where A = 1.50 m>s3 and B = 0.120 m>s4.
The motorcycle is at rest at the origin at time t = 0. (a) Find its
position and velocity as functions of time. (b) Calculate the maximum velocity it attains.
2.54 .. BIO Flying Leap of the Flea. High-speed motion pictures 13500 frames>second2 of a jumping, 210@mg flea yielded the
data used to plot the graph in Fig. E2.54. (See “The Flying Leap
of the Flea” by M. Rothschild, Y. Schlein, K. Parker, C. Neville,
and S. Sternberg in the November 1973 Scientific American.)
This flea was about 2 mm long and jumped at a nearly vertical
takeoff angle. Use the graph to answer these questions: (a) Is the
acceleration of the flea ever zero? If so, when? Justify your answer. (b) Find the maximum height the flea reached in the first
2.5 ms. (c) Find the flea’s acceleration at 0.5 ms, 1.0 ms, and
1.5 ms. (d) Find the flea’s height at 0.5 ms, 1.0 ms, and 1.5 ms.
2.51
proBLems
. BIO A typical male sprinter can maintain his maximum
acceleration for 2.0 s, and his maximum speed is 10 m> s. After
he reaches this maximum speed, his acceleration becomes zero,
and then he runs at constant speed. Assume that his acceleration
is constant during the first 2.0 s of the race, that he starts from
rest, and that he runs in a straight line. (a) How far has the sprinter
run when he reaches his maximum speed? (b) What is the magnitude of his average velocity for a race of these lengths: (i) 50.0 m;
(ii) 100.0 m; (iii) 200.0 m?
2.56 . CALC A lunar lander is descending toward the moon’s
surface. Until the lander reaches the surface, its height above the
surface of the moon is given by y1t2 = b - ct + dt 2, where
b = 800 m is the initial height of the lander above the surface,
c = 60.0 m>s, and d = 1.05 m>s2. (a) What is the initial velocity
of the lander, at t = 0? (b) What is the velocity of the lander just
before it reaches the lunar surface?
2.57 ... Earthquake Analysis. Earthquakes produce several
types of shock waves. The most well known are the P-waves (P for
primary or pressure) and the S-waves (S for secondary or shear).
In the earth’s crust, P-waves travel at about 6.5 km>s and S-waves
move at about 3.5 km>s. The time delay between the arrival of
these two waves at a seismic recording station tells geologists how
far away an earthquake occurred. If the time delay is 33 s, how far
from the seismic station did the earthquake occur?
2.58 .. A brick is dropped from the roof of a tall building. After
it has been falling for a few seconds, it falls 40.0 m in a 1.00-s
time interval. What distance will it fall during the next 1.00 s?
Ignore air resistance.
2.59 ... A rocket carrying a satellite is accelerating straight up
from the earth’s surface. At 1.15 s after liftoff, the rocket clears
the top of its launch platform, 63 m above the ground. After an additional 4.75 s, it is 1.00 km above the ground. Calculate the magnitude of the average velocity of the rocket for (a) the 4.75-s part of
its flight and (b) the first 5.90 s of its flight.
2.60 ... A subway train starts from rest at a station and accelerates at a rate of 1.60 m>s2 for 14.0 s. It runs at constant speed for
70.0 s and slows down at a rate of 3.50 m>s2 until it stops at the
next station. Find the total distance covered.
2.61 . A gazelle is running in a straight line (the x-axis). The
graph in Fig. P2.61 shows this animal’s velocity as a function of
time. During the first 12.0 s, find (a) the total distance moved
and (b) the displacement of the gazelle. (c) Sketch an ax@t graph
showing this gazelle’s acceleration as a function of time for the
first 12.0 s.
2.55
Figure P2.61
vx (m>s)
12.0
8.00
Figure e2.54
4.00
Speed (cm>s)
150
O
100
2.62
50
O
0.5
1.0
1.5
Time (ms)
2.0
2.5
..
2.00 4.00 6.00 8.00 10.0 12.0
t (s)
Collision. The engineer of a passenger train traveling
at 25.0 m>s sights a freight train whose caboose is 200 m ahead on
the same track (Fig. P2.62). The freight train is traveling at
15.0 m>s in the same direction as the passenger train. The engineer of the passenger train immediately applies the brakes,
problems
Figure P2.62
63
... CALC The acceleration of a particle is given by ax1t2 =
-2.00 m>s2 + 13.00 m>s32t. (a) Find the initial velocity v0x such
that the particle will have the same x-coordinate at t = 4.00 s as it
had at t = 0. (b) What will be the velocity at t = 4.00 s ?
2.70 . Egg Drop. You are Figure P2.70
on the roof of the physics building, 46.0 m above the ground
(Fig. P2.70). Your physics professor, who is 1.80 m tall, is
walking alongside the building
at a constant speed of 1.20 m>s.
If you wish to drop an egg on
46.0 m
your professor’s head, where
should the professor be when
you release the egg? Assume
v = 1.20 m>s
that the egg is in free fall.
2.69
vPT = 25.0 m>s
a = -0.100 m>s2
vFT = 15.0 m>s
200 m
causing a constant acceleration of 0.100 m>s2 in a direction opposite to the train’s velocity, while the freight train continues with
constant speed. Take x = 0 at the location of the front of the passenger train when the engineer applies the brakes. (a) Will the
cows nearby witness a collision? (b) If so, where will it take place?
(c) On a single graph, sketch the positions of the front of the passenger train and the back of the freight train.
2.63 ... A ball starts from rest and rolls down a hill with uniform acceleration, traveling 200 m during the second 5.0 s of its
motion. How far did it roll during the first 5.0 s of motion?
2.64 .. Two cars start 200 m apart and drive toward each other
at a steady 10 m>s. On the front of one of them, an energetic grasshopper jumps back and forth between the cars (he has strong
legs!) with a constant horizontal velocity of 15 m>s relative to the
ground. The insect jumps the instant he lands, so he spends no
time resting on either car. What total distance does the grasshopper travel before the cars hit?
2.65 . A car and a truck start from rest at the same instant, with
the car initially at some distance behind the truck. The truck has a
constant acceleration of 2.10 m>s2, and the car has an acceleration
of 3.40 m>s2. The car overtakes the truck after the truck has moved
60.0 m. (a) How much time does it take the car to overtake the
truck? (b) How far was the car behind the truck initially? (c) What
is the speed of each when they are abreast? (d) On a single graph,
sketch the position of each vehicle as a function of time. Take
x = 0 at the initial location of the truck.
2.66 .. You are standing at rest at a bus stop. A bus moving at a
constant speed of 5.00 m>s passes you. When the rear of the bus is
12.0 m past you, you realize that it is your bus, so you start to run
toward it with a constant acceleration of 0.960 m>s2. How far
would you have to run before you catch up with the rear of the bus,
and how fast must you be running then? Would an average college
student be physically able to accomplish this?
2.67 .. Passing. The driver of a car wishes to pass a truck that
is traveling at a constant speed of 20.0 m>s (about 45 mi>h).
Initially, the car is also traveling at 20.0 m>s, and its front bumper
is 24.0 m behind the truck’s rear bumper. The car accelerates at a
constant 0.600 m>s2, then pulls back into the truck’s lane when the
rear of the car is 26.0 m ahead of the front of the truck. The car is
4.5 m long, and the truck is 21.0 m long. (a) How much time is required for the car to pass the truck? (b) What distance does the car
travel during this time? (c) What is the final speed of the car?
2.68 .. CALC An object’s velocity is measured to be vx1t2 =
a - bt 2, where a = 4.00 m>s and b = 2.00 m>s3. At t = 0 the
object is at x = 0. (a) Calculate the object’s position and acceleration as functions of time. (b) What is the object’s maximum
positive displacement from the origin?
1.80 m
. A certain volcano on earth can eject rocks vertically to a
maximum height H. (a) How high (in terms of H) would these rocks
go if a volcano on Mars ejected them with the same initial velocity?
The acceleration due to gravity on Mars is 3.71 m>s2; ignore air resistance on both planets. (b) If the rocks are in the air for a time T on
earth, for how long (in terms of T) would they be in the air on Mars?
2.72 .. An entertainer juggles balls while doing other activities.
In one act, she throws a ball vertically upward, and while it is in
the air, she runs to and from a table 5.50 m away at an average
speed of 3.00 m>s, returning just in time to catch the falling ball.
(a) With what minimum initial speed must she throw the ball
upward to accomplish this feat? (b) How high above its initial
position is the ball just as she reaches the table?
2.73 ... Look Out Below. Sam heaves a 16-lb shot straight up,
giving it a constant upward acceleration from rest of 35.0 m>s2
for 64.0 cm. He releases it 2.20 m above the ground. Ignore air
resistance. (a) What is the speed of the shot when Sam releases it?
(b) How high above the ground does it go? (c) How much time
does he have to get out of its way before it returns to the height of
the top of his head, 1.83 m above the ground?
2.74 ... A flowerpot falls off a windowsill and passes the window of the story below. Ignore air resistance. It takes the pot
0.380 s to pass from the top to the bottom of this window, which is
1.90 m high. How far is the top of the window below the windowsill from which the flowerpot fell?
2.75 .. Two stones are thrown vertically upward from the ground,
one with three times the initial speed of the other. (a) If the faster
stone takes 10 s to return to the ground, how long will it take the
slower stone to return? (b) If the slower stone reaches a maximum
height of H, how high (in terms of H) will the faster stone go?
Assume free fall.
2.76 ... A Multistage Rocket. In the first stage of a two-stage
rocket, the rocket is fired from the launch pad starting from rest
but with a constant acceleration of 3.50 m>s2 upward. At 25.0 s
after launch, the second stage fires for 10.0 s, which boosts the
rocket’s velocity to 132.5 m>s upward at 35.0 s after launch. This
firing uses up all of the fuel, however, so after the second stage
has finished firing, the only force acting on the rocket is gravity.
Ignore air resistance. (a) Find the maximum height that the stagetwo rocket reaches above the launch pad. (b) How much time after
the end of the stage-two firing will it take for the rocket to fall
back to the launch pad? (c) How fast will the stage-two rocket be
moving just as it reaches the launch pad?
2.71
64
ChApTer 2 Motion Along a Straight Line
2.77 ... During your summer internship for an aerospace company, you are asked to design a small research rocket. The rocket
is to be launched from rest from the earth’s surface and is to reach
a maximum height of 960 m above the earth’s surface. The rocket’s engines give the rocket an upward acceleration of 16.0 m>s2
during the time T that they fire. After the engines shut off, the
rocket is in free fall. Ignore air resistance. What must be the value
of T in order for the rocket to reach the required altitude?
2.78 .. A physics teacher performing an outdoor demonstration
suddenly falls from rest off a high cliff and simultaneously shouts
“Help.” When she has fallen for 3.0 s, she hears the echo of her
shout from the valley floor below. The speed of sound is 340 m>s.
(a) How tall is the cliff? (b) If we ignore air resistance, how fast
will she be moving just before she hits the ground? (Her actual
speed will be less than this, due to air resistance.)
2.79 ... A helicopter carrying Dr. Evil takes off with a constant
upward acceleration of 5.0 m>s2. Secret agent Austin Powers
jumps on just as the helicopter lifts off the ground. After the two
men struggle for 10.0 s, Powers shuts off the engine and steps out
of the helicopter. Assume that the helicopter is in free fall after its
engine is shut off, and ignore the effects of air resistance. (a) What
is the maximum height above ground reached by the helicopter?
(b) Powers deploys a jet pack strapped on his back 7.0 s after leaving the helicopter, and then he has a constant downward acceleration with magnitude 2.0 m>s2. How far is Powers above the ground
when the helicopter crashes into the ground?
2.80 .. Cliff Height. You are climbing in the High Sierra when
you suddenly find yourself at the edge of a fog-shrouded cliff. To
find the height of this cliff, you drop a rock from the top; 8.00 s
later you hear the sound of the rock hitting the ground at the foot
of the cliff. (a) If you ignore air resistance, how high is the cliff if
the speed of sound is 330 m>s ? (b) Suppose you had ignored the
time it takes the sound to reach you. In that case, would you have
overestimated or underestimated the height of the cliff? Explain.
2.81 .. CALC An object is moving along the x-axis. At t = 0 it
has velocity v0x = 20.0 m>s. Starting at time t = 0 it has acceleration ax = - Ct, where C has units of m>s3. (a) What is the
value of C if the object stops in 8.00 s after t = 0? (b) For the
value of C calculated in part (a), how far does the object travel
during the 8.00 s?
2.82 .. A ball is thrown straight up from the ground with speed v0.
At the same instant, a second ball is dropped from rest from a
height H, directly above the point where the first ball was thrown
upward. There is no air resistance. (a) Find the time at which the
two balls collide. (b) Find the value of H in terms of v0 and g such
that at the instant when the balls collide, the first ball is at the
highest point of its motion.
2.83 . CALC Cars A and B travel in a straight line. The distance
of A from the starting point is given as a function of time by
xA1t2 = at + bt 2, with a = 2.60 m>s and b = 1.20 m>s2. The
distance of B from the starting point is xB1t2 = gt 2 - dt 3, with
g = 2.80 m>s2 and d = 0.20 m>s3. (a) Which car is ahead just
after the two cars leave the starting point? (b) At what time(s) are
the cars at the same point? (c) At what time(s) is the distance from
A to B neither increasing nor decreasing? (d) At what time(s) do A
and B have the same acceleration?
2.84 .. DATA In your physics lab you release a small glider
from rest at various points on a long, frictionless air track that is
inclined at an angle u above the horizontal. With an electronic
photocell, you measure the time t it takes the glider to slide a
distance x from the release point to the bottom of the track.
Your measurements are given in Fig. P2.84, which shows a
Figure P2.84
t (s)
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.50 1.00 1.50 2.00 2.50 3.00
x (m)
second-order polynomial (quadratic) fit to the plotted data. You
are asked to find the glider’s acceleration, which is assumed to be
constant. There is some error in each measurement, so instead of
using a single set of x and t values, you can be more accurate if
you use graphical methods and obtain your measured value of the
acceleration from the graph. (a) How can you re-graph the data so
that the data points fall close to a straight line? (Hint: You might
want to plot x or t, or both, raised to some power.) (b) Construct
the graph you described in part (a) and find the equation for the
straight line that is the best fit to the data points. (c) Use the straightline fit from part (b) to calculate the acceleration of the glider.
(d) The glider is released at a distance x = 1.35 m from the
bottom of the track. Use the acceleration value you obtained in
part (c) to calculate the speed of the glider when it reaches the bottom of the track.
2.85 .. DATA In a physics lab experiment, you release a small
steel ball at various heights above the ground and measure the
ball’s speed just before it strikes the ground. You plot your data on
a graph that has the release height (in meters) on the vertical axis
and the square of the final speed (in m2>s2) on the horizontal axis.
In this graph your data points lie close to a straight line. (a) Using
g = 9.80 m>s2 and ignoring the effect of air resistance, what is
the numerical value of the slope of this straight line? (Include the
correct units.) The presence of air resistance reduces the magnitude of the downward acceleration, and the effect of air resistance
increases as the speed of the object increases. You repeat the experiment, but this time with a tennis ball as the object being dropped.
Air resistance now has a noticeable effect on the data. (b) Is the
final speed for a given release height higher than, lower than, or
the same as when you ignored air resistance? (c) Is the graph of
the release height versus the square of the final speed still a
straight line? Sketch the qualitative shape of the graph when air
resistance is present.
2.86 ... DATA A model car starts from rest and travels in a
straight line. A smartphone mounted on the car has an app that
transmits the magnitude of the car’s acceleration (measured by an
accelerometer) every second. The results are given in the table:
Time (s)
0
1.00
2.00
3.00
4.00
5.00
Acceleration 1m , s2 2
5.95
5.52
5.08
4.55
3.96
3.40
passage problems
chaLLenge probLems
2.87
...
In the vertical jump, an athlete starts from a crouch and
jumps upward as high as possible. Even the best athletes spend
little more than 1.00 s in the air (their “hang time”). Treat the athlete as a particle and let ymax be his maximum height above the
floor. To explain why he seems to hang in the air, calculate the
ratio of the time he is above ymax>2 to the time it takes him to go
from the floor to that height. Ignore air resistance.
2.88 ... Catching the Bus. A student is running at her top
speed of 5.0 m>s to catch a bus, which is stopped at the bus stop.
When the student is still 40.0 m from the bus, it starts to pull
away, moving with a constant acceleration of 0.170 m>s2. (a) For
how much time and what distance does the student have to run at
5.0 m>s before she overtakes the bus? (b) When she reaches the
bus, how fast is the bus traveling? (c) Sketch an x-t graph for both
the student and the bus. Take x = 0 at the initial position of the
student. (d) The equations you used in part (a) to find the time
have a second solution, corresponding to a later time for which the
student and bus are again at the same place if they continue their
specified motions. Explain the significance of this second solution. How fast is the bus traveling at this point? (e) If the student’s
top speed is 3.5 m>s, will she catch the bus? (f) What is the minimum speed the student must have to just catch up with the bus?
For what time and what distance does she have to run in that case?
2.89 ... A ball is thrown straight up from the edge of the roof of
a building. A second ball is dropped from the roof 1.00 s later.
Ignore air resistance. (a) If the height of the building is 20.0 m,
what must the initial speed of the first ball be if both are to hit the
ground at the same time? On the same graph, sketch the positions
of both balls as a function of time, measured from when the first
ball is thrown. Consider the same situation, but now let the initial
speed v0 of the first ball be given and treat the height h of the
building as an unknown. (b) What must the height of the building
be for both balls to reach the ground at the same time if (i) v0 is
6.0 m>s and (ii) v0 is 9.5 m>s ? (c) If v0 is greater than some value
vmax, no value of h exists that allows both balls to hit the ground at
the same time. Solve for vmax. The value vmax has a simple physical
interpretation. What is it? (d) If v0 is less than some value vmin, no
value of h exists that allows both balls to hit the ground at the
same time. Solve for vmin. The value vmin also has a simple physical
interpretation. What is it?
passage probLems
BIO Blood Flow in ThE hEarT. The human circulatory
system is closed—that is, the blood pumped out of the left ventricle of the heart into the arteries is constrained to a series of continuous, branching vessels as it passes through the capillaries and
then into the veins as it returns to the heart. The blood in each of
the heart’s four chambers comes briefly to rest before it is ejected
by contraction of the heart muscle.
2.90 If the contraction of the left ventricle lasts 250 ms and the
speed of blood flow in the aorta (the large artery leaving the heart)
is 0.80 m>s at the end of the contraction, what is the average
acceleration of a red blood cell as it leaves the heart? (a) 310 m>s2;
(b) 31 m>s2; (c) 3.2 m>s2; (d) 0.32 m>s2.
2.91 If the aorta (diameter da) branches into two equal-sized
arteries with a combined area equal to that of the aorta, what is
the diameter of one of the branches? (a) 1da; (b) da >12; (c) 2da;
(d) da >2.
2.92 The velocity of blood in the aorta can be measured directly
with ultrasound techniques. A typical graph of blood velocity versus time during a single heartbeat is shown in Fig. P2.92. Which
statement is the best interpretation of this graph? (a) The blood
flow changes direction at about 0.25 s; (b) the speed of the blood
flow begins to decrease at about 0.10 s; (c) the acceleration of the
blood is greatest in magnitude at about 0.25 s; (d) the acceleration
of the blood is greatest in magnitude at about 0.10 s.
Figure P2.92
Blood velocity (m>s)
Each measured value has some experimental error. (a) Plot acceleration versus time and find the equation for the straight line that
gives the best fit to the data. (b) Use the equation for a1t2 that you
found in part (a) to calculate v1t2, the speed of the car as a function of time. Sketch the graph of v versus t. Is this graph a straight
line? (c) Use your result from part (b) to calculate the speed of the
car at t = 5.00 s. (d) Calculate the distance the car travels between t = 0 and t = 5.00 s.
65
1.0
0.8
0.6
0.4
0.2
0
0.25
0.50
0.75
Time (s)
1.00
66
Chapter 2 Motion along a Straight Line
answers
chapter opening Question
?
(iii) Acceleration refers to any change in velocity, including both
speeding up and slowing down.
test your understanding Questions
2.1 (a): (iv), (i) and (iii) (tie), (v), (ii); (b): (i) and (iii); (c): (v) In
(a) the average x-velocity is vav­x = ∆x>∆t. For all five trips, ∆t =
1 h. For the individual trips, (i) ∆x = + 50 km, vav­x = + 50 km>h;
(ii) ∆x = -50 km, vav­x = - 50 km>h; (iii) ∆x = 60 km - 10 km =
+ 50 km, vav­x = + 50 km>h; (iv) ∆x = + 70 km, vav­x = +70 km>h;
(v) ∆x = - 20 km + 20 km = 0, vav­x = 0. In (b) both have vav­x =
+ 50 km>h.
2.2 (a) P, Q and S (tie), R The x-velocity is (b) positive when
the slope of the x-t graph is positive (P), (c) negative when the
slope is negative (R), and (d) zero when the slope is zero (Q and S).
(e) R, P, Q and S (tie) The speed is greatest when the slope of the
x-t graph is steepest (either positive or negative) and zero when
the slope is zero.
2.3 (a) S, where the x-t graph is curved upward (concave up).
(b) Q, where the x-t graph is curved downward (concave down).
(c) P and R, where the x-t graph is not curved either up or down.
(d) At P, ax = 0 (velocity is not changing); at Q, ax 6 0 (velocity is decreasing, i.e., changing from positive to zero to negative);
at R, ax = 0 (velocity is not changing); and at S, ax 7 0 (velocity
is increasing, i.e., changing from negative to zero to positive).
2.4 (b) The officer’s x-acceleration is constant, so her vx@t
graph is a straight line. The motorcycle is moving faster than the
car when the two vehicles meet at t = 10 s.
2.5 (a) (iii) Use Eq. (2.13) with x replaced by y and ay = - g;
v y2 = v 0y 2 - 2g1y - y02. The starting height is y0 = 0 and
the y-velocity at the maximum height y = h is vy = 0, so 0 =
v 0y 2 - 2gh and h = v 0y 2>2g. If the initial y-velocity is increased
by a factor of 2, the maximum height increases by a factor of
22 = 4 and the ball goes to height 4h. (b) (v) Use Eq. (2.8) with x
replaced by y and ay = -g; vy = v0y - gt. The y-velocity at the
maximum height is vy = 0, so 0 = v0y - gt and t = v0y>g. If the
initial y-velocity is increased by a factor of 2, the time to reach
the maximum height increases by a factor of 2 and becomes 2t.
2.6 (ii) The acceleration ax is equal to the slope of the vx@t graph.
If ax is increasing, the slope of the vx­t graph is also increasing
and the graph is concave up.
bridging problem
h = 57.1 m
?
If a cyclist is going around
a curve at constant speed,
is he accelerating? If so, what
is the direction of his acceleration? (i) No; (ii) yes, in the
direction of his motion; (iii) yes,
toward the inside of the curve;
(iv) yes, toward the outside of
the curve; (v) yes, but in some
other direction.
3
Motion in two or
three DiMensions
Learning goaLs
Looking forward at …
3.1 How to use vectors to represent the position
3.2
3.3
3.4
3.5
and velocity of a particle in two or three
dimensions.
How to find the vector acceleration of a
particle, why a particle can have an acceleration even if its speed is constant, and how
to interpret the components of acceleration
parallel and perpendicular to a particle’s
path.
How to solve problems that involve the
curved path followed by a projectile.
How to analyze motion in a circular path,
with either constant speed or varying
speed.
How to relate the velocities of a moving
body as seen from two different frames of
reference.
W
hat determines where a batted baseball lands? How do you describe
the motion of a roller coaster car along a curved track or the flight of a
circling hawk? Which hits the ground first: a baseball that you simply
drop or one that you throw horizontally?
We can’t answer these kinds of questions by using the techniques of Chapter 2,
in which particles moved only along a straight line. Instead, we need to extend
our descriptions of motion to two- and three-dimensional situations. We’ll still
use the vector quantities displacement, velocity, and acceleration, but now these
quantities will no longer lie along a single line. We’ll find that several important
kinds of motion take place in two dimensions only—that is, in a plane.
We also need to consider how the motion of a particle is described by different
observers who are moving relative to each other. The concept of relative velocity
will play an important role later in the book when we explore electromagnetic
phenomena and when we introduce Einstein’s special theory of relativity.
This chapter merges the vector mathematics of Chapter 1 with the kinematic
language of Chapter 2. As before, we’re concerned with describing motion, not
with analyzing its causes. But the language you learn here will be an essential
tool in later chapters when we study the relationship between force and motion.
Looking back at …
2.1 Average x-velocity.
2.2 Instantaneous x-velocity.
2.3 Average and instantaneous x-acceleration.
2.4 Straight-line motion with constant
acceleration.
2.5 The motion of freely falling bodies.
3.1 Position and VeLocity Vectors
Let’s see how to describe a particle’s motion in space. If the particle is at a point P
S
at a certain instant, the position vector r of the particle at this instant is a vector
that goes from the origin of the coordinate system to point P (Fig. 3.1 on next page).
The Cartesian coordinates x, y, and z of point P are the x-, y-, and z-components
S
of vector r . Using the unit vectors we introduced in Section 1.9, we can write
Position vector of a
particle at a given instant
n
r = xdn + yen + zk
S
Unit vectors in x-, y-, and z-directions
Coordinates of
particle’s position
(3.1)
67
68
Chapter 3 Motion in two or three Dimensions
S
3.1 The position vector r from origin O to
point P has components x, y, and z.
y
Position P of a particle
at a given time has
coordinates x, y, z.
y
During a time interval ∆t the particle moves from P1, where its position
S
S
vector is r1, to P2, where its position vector is r2. The change in position (the disS
S
S
placement) during this interval is ∆r = r2 − r1 = 1x2 - x12dn + 1y2 - y12en +
S
1z2 - z12kn . We define the average velocity vav during this interval in the same
way we did in Chapter 2 for straight-line motion, as the displacement divided by
the time interval (Fig. 3.2):
z nk
S
r
O
yen
x
xdn
z
Change in the particle’s position vector
P
Average velocity vector
of a particle during time
interval from t1 to t2
x
Time interval
z
Position vector of point P
has components x, y, z:
S
r = xdn + yen + z nk.
S
3.2 The average velocity vav between
points P1 and P2 has the same direction as
S
the displacement ∆r .
Position at time t2
y
P2
S
r2
S
vav =
S
Displacement
S
vector ∆ r points
from P1 to P2.
∆r
S
r1
O
P1
Position at time t1
x
Particle’s path
z
∆r
∆t
S
S
3.3 The vectors v1 and v2 are the instan-
taneous velocities at the points P1 and P2
shown in Fig. 3.2.
S
Final position
minus initial
position
(3.2)
Final time minus initial time
Dividing a vector by a scalar is a special case of multiplying a vector by a
S
scalar, described in Section 1.7; the average velocity vav is equal to the displaceS
ment vector ∆r multiplied by 1>∆t. Note that the x-component of Eq. (3.2) is
vav - x = 1x2 - x12>1t2 - t12 = ∆x>∆t. This is just Eq. (2.2), the expression for
average x-velocity that we found in Section 2.1 for one-dimensional motion.
We now define instantaneous velocity just as we did in Chapter 2: It equals
the instantaneous rate of change of position with time. The key difference is that
S
S
both position r and instantaneous velocity v are now vectors:
S
The instantaneous velocity
vector of a particle ...
S
S
S
r2 − r1
∆r
=
∆t
t2 - t1
S
vav =
S
dr
∆r
v = lim
=
∆t S 0 ∆t
dt
S
... equals the limit of its average velocity
vector as the time interval approaches zero ...
(3.3)
... and equals the instantaneous rate
of change of its position vector.
S
At any instant, the magnitude of v is the speed v of the particle at that instant, and
S
the direction of v is the direction in which the particle is moving at that instant.
As ∆t S 0, points P1 and P2 in Fig. 3.2 move closer and closer together. In
S
S
this limit, the vector ∆r becomes tangent to the path. The direction of ∆r in this
S
limit is also the direction of v. So at every point along the path, the instantaneous velocity vector is tangent to the path at that point (Fig. 3.3).
It’s often easiest to calculate the instantaneous velocity vector by using
S
components. During any displacement ∆r , the changes ∆x, ∆y, and ∆z in the
S
three coordinates of the particle are the components of ∆r . It follows that the
S
components vx , vy , and vz of the instantaneous velocity v = vxnd + vyne + vz kn are
simply the time derivatives of the coordinates x, y, and z:
y
Each component of a particle’s instantaneous velocity vector ...
S
v2
vx =
P2
z
Particle’s path
vy =
dy
dt
vz =
dz
dt
(3.4)
... equals the instantaneous rate of change of its corresponding coordinate.
The instantaneous
S
velocity vector v is
tangent to the path
at each point.
O
dx
dt
S
S
v1
P1
x
The x-component of v is vx = dx>dt, which is the same as Eq. (2.3) for straightline motion (see Section 2.2). Hence Eq. (3.4) is a direct extension of instantaneous velocity to motion in three dimensions.
We can also get Eq. (3.4) by taking the derivative of Eq. (3.1). The unit vectors
nd , ne , and kn don’t depend on time, so their derivatives are zero and we find
S
v=
S
dy
dr
dx
dz
= nd + ne + kn
dt
dt
dt
dt
(3.5)
S
This shows again that the components of v are dx>dt, dy>dt, and dz>dt.
S
The magnitude of the instantaneous velocity vector v—that is, the speed—is
given in terms of the components vx , vy , and vz by the Pythagorean relation:
0 vS 0 = v = 2v x 2 + v y 2 + v z 2
(3.6)
69
3.1 Position and Velocity Vectors
Figure 3.4 shows the situation when the particle moves in the xy-plane. In this
S
case, z and vz are zero. Then the speed (the magnitude of v) is
3.4 The two velocity components for
motion in the xy-plane.
S
v = 2v x 2 + v y 2
y
S
and the direction of the instantaneous velocity v is given by angle a (the Greek
letter alpha) in the figure. We see that
tan a =
vy
The instantaneous velocity vector v
is always tangent to the path.
S
v
vy
Particle’s path in
the xy-plane
(3.7)
vx
a
vx
(We use a for the direction of the instantaneous velocity vector to avoid confusion with the direction u of the position vector of the particle.)
From now on, when we use the word “velocity,” we will always mean the
S
instantaneous velocity vector v (rather than the average velocity vector). Usually,
S
we won’t even bother to call v a vector; it’s up to you to remember that velocity is
a vector quantity with both magnitude and direction.
O
vx and vy are the x- and yS
components of v.
x
Solution
ExamplE 3.1 CalCulating avEragE and instantanEous vEloCity
A robotic vehicle, or rover, is exploring the surface of Mars. The
stationary Mars lander is the origin of coordinates, and the surrounding Martian surface lies in the xy-plane. The rover, which we
represent as a point, has x- and y-coordinates that vary with time:
x = 2.0 m - 10.25 m > s22t 2
y = 11.0 m > s2t + 10.025 m > s32t 3
(a) Find the rover’s coordinates and distance from the lander at
t = 2.0 s. (b) Find the rover’s displacement and average velocity
vectors for the interval t = 0.0 s to t = 2.0 s. (c) Find a general
S
expression for the rover’s instantaneous velocity vector v. Express
S
v at t = 2.0 s in component form and in terms of magnitude and
direction.
execute: (a) At t = 2.0 s the rover’s coordinates are
x = 2.0 m - 10.25 m > s2212.0 s22 = 1.0 m
y = 11.0 m > s212.0 s2 + 10.025 m > s3212.0 s23 = 2.2 m
The rover’s distance from the origin at this time is
r = 2x 2 + y2 = 2(1.0 m)2 + 12.2 m22 = 2.4 m
(b) To find the displacement and average velocity over the given
S
time interval, we first express the position vector r as a function of
time t. From Eq. (3.1) this is
S
r = xdn + yen
= 32.0 m - 10.25 m > s22t 24 nd
soLution
+ 311.0 m>s2t + 10.025 m>s32t 34 ne
identify and set uP: This problem involves motion in two di-
mensions, so we must use the vector equations obtained in this
section. Figure 3.5 shows the rover’s path (dashed line). We’ll use
S
S
S
S
Eq. (3.1) for position r , the expression ∆r = r2 − r1 for displacement, Eq. (3.2) for average velocity, and Eqs. (3.5), (3.6), and (3.7)
for instantaneous velocity and its magnitude and direction.
S
At t = 0.0 s the position vector r0 is
S
r0 = 12.0 m2dn + 10.0 m2en
S
From part (a), the position vector r2 at t = 2.0 s is
S
r2 = 11.0 m2dn + 12.2 m2en
S
3.5 At t = 0.0 s the rover has position vector r0 and instantaneous
S
S
S
velocity vector v0 . Likewise, r1 and v1 are the vectors at
S
S
t = 1.0 s; r2 and v2 are the vectors at t = 2.0 s.
y (m)
2.0
1.5
a = 128°
S
v1
r2
t = 1.0 s
1.0
S
Rover’s path
S
v0
t = 0.0 s
0.5
S
r0
O
0.5
1.0
1.5
S
vav =
S
r1
S
During this interval the rover moves 1.0 m in the negative
x-direction and 2.2 m in the positive y-direction. From Eq. (3.2),
the average velocity over this interval is the displacement divided
by the elapsed time:
t = 2.0 s
S
S
∆r = r2 − r0 = 11.0 m2dn + 12.2 m2en − 12.0 m2dn
= 1-1.0 m2dn + 12.2 m2en
S
v2
2.5
The displacement from t = 0.0 s to t = 2.0 s is therefore
2.0
x (m)
S
1-1.0 m2dn + 12.2 m2en
∆r
=
2.0 s - 0.0 s
∆t
= 1-0.50 m > s2dn + 11.1 m > s2en
The components of this average velocity are vav@x = - 0.50 m >s
and vav@y = 1.1 m > s.
Continued
70
Chapter 3 Motion in two or three Dimensions
(c) From Eq. (3.4) the components of instantaneous velocity
are the time derivatives of the coordinates:
vx =
vy =
dx
= 1- 0.25 m > s2212t2
dt
S
Figure 3.5 shows the direction of velocity vector v2, which is at an
angle a between 90° and 180° with respect to the positive x-axis.
From Eq. (3.7) we have
arctan
dy
= 1.0 m > s + 10.025 m > s3213t 22
dt
vy
vx
= arctan
1.3 m > s
= -52°
-1.0 m > s
This is off by 180°; the correct value is a = 180° - 52° = 128°,
or 38° west of north.
Hence the instantaneous velocity vector is
S
v = vxnd + vy ne
= 1- 0.50 m > s22tdn + 31.0 m>s + 10.075 m > s32t 24 ne
S
At t = 2.0 s the velocity vector v2 has components
v2x = 1- 0.50 m > s2212.0 s2 = - 1.0 m > s
v2y = 1 .0 m > s + 10.075 m > s3212.0 s22 = 1.3 m > s
The magnitude of the instantaneous velocity (that is, the speed) at
t = 2.0 s is
v2 = 2v 2x2 + v 2y2 = 21 - 1.0 m> s22 + 11.3 m > s22
= 1.6 m > s
eValuaTe: Compare the components of average velocity from
part (b) for the interval from t = 0.0 s to t = 2.0 s 1vav@x =
-0.50 m > s, vav@y = 1.1 m > s2 with the components of instantaneous velocity at t = 2.0 s from part (c) 1v2x = - 1.0 m > s, v2y =
S
1.3 m > s2. Just as in one dimension, the average velocity vector vav
over an interval is in general not equal to the instantaneous
S
velocity v at the end of the interval (see Example 2.1).
S
Figure 3.5 shows the position vectors r and instantaneous velocS
ity vectors v at t = 0.0 s, 1.0 s, and 2.0 s. (Calculate these quantiS
ties for t = 0.0 s and t = 1.0 s.) Notice that v is tangent to the path
S
at every point. The magnitude of v increases as the rover moves,
which means that its speed is increasing.
TesT Your undersTanding of secTion 3.1 In which of these situations
S
would the average velocity vector vav over an interval be equal to the instantaneous
S
velocity v at the end of the interval? (i) A body moving along a curved path at constant
speed; (ii) a body moving along a curved path and speeding up; (iii) a body moving
along a straight line at constant speed; (iv) a body moving along a straight line and
speeding up. ❙
3.2 The acceleraTion VecTor
Now let’s consider the acceleration of a particle moving in space. Just as for
motion in a straight line, acceleration describes how the velocity of the particle
changes. But since we now treat velocity as a vector, acceleration will describe
changes in the velocity magnitude (that is, the speed) and changes in the direction of velocity (that is, the direction in which the particle is moving).
In Fig. 3.6a, a car (treated as a particle) is moving along a curved road. Vectors
S
S
v1 and v2 represent the car’s instantaneous velocities at time t1, when the car is
S
S
S
3.6 (a) A car moving along a curved road from P1 to P2 . (b) How to obtain the change in velocity ∆v = v2 − v1
S
S
by vector subtraction. (c) The vector aav = ∆v/∆t represents the average acceleration between P1 and P2 .
(b)
(a)
(c)
S
v2
S
v1
P1
S
v2
P2
This car accelerates by slowing
while rounding a curve. (Its
instantaneous velocity changes in
both magnitude and direction.)
S
v2
P2
P2
S
v1
S
S
v1
P1
S
S
v1
S
∆ v = v2 − v1
S
v2
To find the car’s average acceleration between
P1 and P2, we first find the change in velocity
S
S
S
∆v by subtracting v1 from v2. (Notice that
S
S
S
v1 + ∆v = v2.)
P1
S
∆v
S
aav =
S
∆v
∆t
The average acceleration has the same direction
S
as the change in velocity, ∆v.
3.2 The Acceleration Vector
71
at point P1, and at time t2, when the car is at point P2 . During the time interval
S
S
S
S
S
S
from t1 to t2, the vector change in velocity is v2 − v1 = ∆v, so v2 = v1 + ∆v
S
(Fig. 3.6b). The average acceleration aav of the car during this time interval is
the velocity change divided by the time interval t2 - t1 = ∆t:
Change in the particle’s velocity
Average acceleration
S
S
S
Final velocity
∆v v2 − v1
vector of a particle
S
minus initial
a
=
=
av
during time interval
∆t
t2 - t1
velocity
from t1 to t2
Time interval
Final time minus initial time
Demo
(3.8)
S
3.7 (a) Instantaneous acceleration a at
point P1 in Fig. 3.6. (b) Instantaneous ac-
S
Average acceleration is a vector quantity in the same direction as ∆v (Fig. 3.6c).
The x-component of Eq. (3.8) is aav - x = 1v2x - v1x2>1t2 - t12 = ∆vx>∆t,
which is just Eq. (2.4) for average acceleration in straight-line motion.
S
As in Chapter 2, we define the instantaneous acceleration a (a vector quantity) at point P1 as the limit of the average acceleration vector when point P2
S
approaches point P1, so both ∆v and ∆t approach zero (Fig. 3.7):
The instantaneous
acceleration vector
of a particle ...
S
S
celeration for motion along a straight line.
(a) Acceleration: curved trajectory
S
To find the instantaneous v2
acceleration
P2
S
S
a at P1 ...
v1
S
... we take the limit of aav
as P2 approaches P1 ...
P1
S
dv
∆v
=
∆t S 0 ∆t
dt
a = lim
(3.9)
S
... equals the limit of its average acceleration
vector as the time interval approaches zero ...
v1
... and equals the instantaneous rate
of change of its velocity vector.
S
The velocity vector v is always tangent to the particle’s path, but the instantaS
neous acceleration vector a does not have to be tangent to the path. If the path is
S
curved, a points toward the concave side of the path—that is, toward the inside
of any turn that the particle is making (Fig. 3.7a). The acceleration is tangent to
the path only if the particle moves in a straight line (Fig. 3.7b).
P1
S
... meaning that ∆v and ∆t
approach 0.
S
S
a = lim ∆v
∆tS0 ∆t
Acceleration points to
concave side of path.
(b) Acceleration: straight-line trajectory
caution Any particle following a curved path is accelerating When a particle is moving in a curved path, it always has nonzero acceleration, even when it moves with constant
speed. This conclusion is contrary to the everyday use of the word “acceleration” to
mean that speed is increasing. The more precise definition given in Eq. (3.9) shows that
there is a nonzero acceleration whenever the velocity vector changes in any way,
whether there is a change of speed, direction, or both. ❙
?
To convince yourself that a particle is accelerating as it moves on a
curved path with constant speed, think of your sensations when you ride in a
car. When the car accelerates, you tend to move inside the car in a direction opposite to the car’s acceleration. (In Chapter 4 we’ll learn why this is so.) Thus you
tend to slide toward the back of the car when it accelerates forward (speeds up)
and toward the front of the car when it accelerates backward (slows down). If the
car makes a turn on a level road, you tend to slide toward the outside of the turn;
hence the car is accelerating toward the inside of the turn.
We’ll usually be interested in instantaneous acceleration, not average acceleration. From now on, we’ll use the term “acceleration” to mean the instantaneous
S
acceleration vector a.
S
Each component of the acceleration vector a = axnd + ayne + az kn is the
derivative of the corresponding component of velocity:
Each component of a particle’s instantaneous acceleration vector ...
ax =
dvx
dt
ay =
dvy
dt
az =
dvz
dt
... equals the instantaneous rate of change of its corresponding velocity component.
(3.10)
Only if the trajectory is
a straight line ...
S
v2
P2
S
v1
P1
S
∆v
S
a = lim
∆tS0
S
∆v
∆t
... is the acceleration
tangent to the trajectory.
BIO application Horses on a Curved
Path By leaning to the side and hitting the
ground with their hooves at an angle, these
horses give themselves the sideways
acceleration necessary to make a sharp
change in direction.
72
ChAPTer 3 Motion in Two or Three Dimensions
In terms of unit vectors,
3.8 When the fingers release the arrow, its
acceleration vector has a horizontal component 1ax2 and a vertical component 1ay2.
S
a=
S
ay
a
ax
ay =
dt 2
d 2y
az =
dt 2
d 2z
(3.12)
dt 2
Solution
Let’s return to the motions of the Mars rover in Example 3.1.
(a) Find the components of the average acceleration for the interval t = 0.0 s to t = 2.0 s. (b) Find the instantaneous acceleration
at t = 2.0 s.
soLution
identify and set uP: In Example 3.1 we found the components
of the rover’s instantaneous velocity at any time t:
dx
= 1- 0.25 m>s2212t2 = 1- 0.50 m>s22t
dt
dy
vy =
= 1.0 m>s + 10.025 m>s3213t 22
dt
= 1.0 m>s + 10.075 m>s32t 2
We’ll use the vector relationships among velocity, average acceleration, and instantaneous acceleration. In part (a) we determine
the values of vx and vy at the beginning and end of the interval and
then use Eq. (3.8) to calculate the components of the average
acceleration. In part (b) we obtain expressions for the instantaneous acceleration components at any time t by taking the time
derivatives of the velocity components as in Eqs. (3.10).
execute: (a) In Example 3.1 we found that at t = 0.0 s the velocity
components are
vx = 0.0 m>s
vy = 1.0 m>s
vx = - 1.0 m>s
S
Hence the instantaneous acceleration vector a at time t is
a = axnd + ay ne = 1-0.50 m>s22dn + 10.15 m>s32ten
S
At t = 2.0 s the components of acceleration and the acceleration
vector are
ax = -0.50 m>s2
S
a = 1- 0.50 m>s22dn + 10.30 m>s22en
The magnitude of acceleration at this time is
a = 2ax2 + ay2
= 21- 0.50 m>s222 + 10.30 m>s222 = 0.58 m>s2
A sketch of this vector (Fig. 3.9) shows that the direction angle b
S
of a with respect to the positive x-axis is between 90° and 180°.
From Eq. (3.7) we have
arctan
vy = 1.3 m>s
-0.50 m>s2
S
2.0
∆vy
1.0
S
v2
ay =
dt
S
a = 128°
b = 149°
S
a2
t = 2.0 s
S
v1
S
a1
t = 1.0 s
Rover’s path
S
v0
0.5
(b) Using Eqs. (3.10), we find
dvy
= - 31°
t = 0.0 s
S
a0
3
= 10.075 m>s 212t2
S
acceleration at t = 0.0 s 1v0 and a02, t = 1.0 s 1v1 and a12, and
S
S
t = 2.0 s 1v2 and a22.
2.5
1.3 m>s - 1.0 m>s
= 0.15 m>s2
2.0 s - 0.0 s
0.30 m>s2
3.9 The path of the robotic rover, showing the velocity and
1.5
dvx
ax =
= - 0.50 m>s2
dt
ax
= arctan
Hence b = 180° + 1-31°2 = 149°.
∆vx
-1.0 m>s - 0.0 m>s
=
= - 0.50 m>s2
∆t
2.0 s - 0.0 s
=
ay
y (m)
Thus the components of average acceleration in the interval
t = 0.0 s to t = 2.0 s are
∆t
ay = 10.15 m>s3212.0 s2 = 0.30 m>s2
S
and that at t = 2.0 s the components are
aav@y =
d 2x
CalCulating avEragE and instantanEous aCCElEration
ExamplE 3.2
aav@x =
(3.11)
The x-component of Eqs. (3.10) and (3.11), ax = dvx>dt, is just Eq. (2.5) for instantaneous acceleration in one dimension. Figure 3.8 shows an example of an
acceleration vector that has both x- and y-components.
Since each component of velocity is the derivative of the corresponding
coordinate, we can express the components ax , ay , and az of the acceleration
S
vector a as
ax =
vx =
S
dvy
dvz n
dvx
dv
nd +
ne +
=
k
dt
dt
dt
dt
O
0.5
1.0
1.5
2.0
x (m)
3.2 the acceleration Vector
73
S
direction at any of these times. The velocity vector v is tangent to
the path at each point (as is always the case), and the acceleration
S
vector a points toward the concave side of the path.
eVaLuate: Figure 3.9 shows the rover’s path and the velocity and
acceleration vectors at t = 0.0 s, 1.0 s, and 2.0 s. (Use the results
of part (b) to calculate the instantaneous acceleration at t = 0.0 s
S
S
and t = 1.0 s for yourself.) Note that v and a are not in the same
Parallel and Perpendicular components of
acceleration
Equations (3.10) tell us about the components of a particle’s instantaneous acS
celeration vector a along the x-, y-, and z-axes. Another useful way to think about
S
a is in terms of one component parallel to the particle’s path and to its velocS
S
ity v, and one component perpendicular to the path and to v (Fig. 3.10). That’s
because the parallel component aŒ tells us about changes in the particle’s speed,
while the perpendicular component a# tells us about changes in the particle’s
S
direction of motion. To see why the parallel and perpendicular components of a
have these properties, let’s consider two special cases.
S
In Fig. 3.11a the acceleration vector is in the same direction as the velocity v1,
S
S
so a has only a parallel component aŒ (that is, a# = 0). The velocity change ∆v
S
during a small time interval ∆t is in the same direction as a and hence in the
S
S
same direction as v1 . The velocity v2 at the end of ∆t is in the same direction
S
as v1 but has greater magnitude. Hence during the time interval ∆t the particle
in Fig. 3.11a moved in a straight line with increasing speed (compare Fig. 3.7b).
S
In Fig. 3.11b the acceleration is perpendicular to the velocity, so a has only
a perpendicular component a# (that is, aŒ = 0). In a small time interval ∆t, the
S
S
S
S
velocity change ∆v is very nearly perpendicular to v1, and so v1 and v2 have
different directions. As the time interval ∆t approaches zero, the angle f in the
S
S
S
S
figure also approaches zero, ∆v becomes perpendicular to both v1 and v2, and v1
S
and v2 have the same magnitude. In other words, the speed of the particle stays
the same, but the direction of motion changes and the path of the particle curves.
S
In the most general case, the acceleration a has both components parallel and
S
perpendicular to the velocity v, as in Fig. 3.10. Then the particle’s speed will
change (described by the parallel component aΠ) and its direction of motion will
change (described by the perpendicular component
a#).parallel to velocity
(a) Acceleration
Figure 3.12 shows a particle moving along a curved path for three situations:
Changes only
magnitude
S
constant speed, increasing speed, and decreasing
speed.
If the speed
∆vis constant,
of velocity:
S
S speed changes;
a is perpendicular, or normal, to the path and
to vdoesn’t.
and points toward
the concave
S
direction
v1
side of the path (Fig. 3.12a). If the speed is increasing, there
is
still
a perpenS
a
S
S
S
dicular component of a, but there is also a parallel component with
thevS 1 same
v2 =
+ ∆v
S
S
direction as v (Fig. 3.12b). Then a points ahead of the normal to the path. (This
was the case in Example 3.2.) If the speed is decreasing, the parallel compoS
S
nent has the direction opposite to v, and a points behind the normal to the
path (Fig. 3.12c; compare Fig. 3.7a). We will use these ideas again in Section 3.4
when we study the special case of motion in a circle.
3.10 The acceleration can be resolved
into a component aΠparallel to the path
(that is, along the tangent to the path) and
a component a# perpendicular to the path
(that is, along the normal to the path).
Tangent to path at P
S
Component of
S
a parallel to
the path
aŒ
v
Particle’s path
S
a
P
a#
S
Component of a
perpendicular to the path
Normal to
path at P
3.11 The effect of acceleration directed
(a) parallel to and (b) perpendicular to a
particle’s velocity.
(a) Acceleration parallel to velocity
(b) Acceler
Changes only magnitude
S
∆v
of velocity: speed changes;
S
direction doesn’t.
v1
Changes on
velocity: pa
curved path
speed.
S
a
S
S
S
v2 = v1 + ∆v
(b) Acceleration perpendicular to velocity
Changes only direction of
velocity: particle follows
curved path at constant
speed.
S
∆v
S
v1
f
S
S
S
v2 = v1 + ∆v
S
a
PhET: Maze Game
3.12 Velocity and acceleration vectors for a particle moving through a point P on
a curved path with (a) constant speed, (b) increasing speed, and (c) decreasing speed.
(a) When speed is constant along a curved
path ...
(b) When speed is increasing along a curved
path ...
S
S
v
P
(c) When speed is decreasing along a curved
path ...
S
v
... acceleration is
normal to the path.
P
v
... acceleration points
ahead of the normal.
P
... acceleration points
behind the normal.
S
a
S
a
Normal at P
S
a
Normal at P
Normal at P
74
Chapter 3 Motion in two or three Dimensions
Solution
ExamplE 3.3
CalCulating parallEl and pErpEndiCular ComponEnts of aCCElEration
For the rover of Examples 3.1 and 3.2, find the parallel and perpendicular components of the acceleration at t = 2.0 s.
3.13 The parallel and perpendicular components of the acceleration of the rover at t = 2.0 s.
S
v
soLution
identify and set uP: We want to find the components of the acS
celeration vector a that are parallel and perpendicular to velocity
S
S
S
vector v. We found the directions of v and a in Examples 3.1 and
3.2, respectively; Fig. 3.9 shows the results. From these directions
we can find the angle between the two vectors and the compoS
S
nents of a with respect to the direction of v.
execute: From Example 3.2, at t = 2.0 s the particle has an acceleration of magnitude 0.58 m>s2 at an angle of 149° with respect
to the positive x-axis. In Example 3.1 we found that at this time the
velocity vector is at an angle of 128° with respect to the positive
S
S
x-axis. The angle between a and v is therefore 149° - 128° = 21°
(Fig. 3.13). Hence the components of acceleration parallel and
S
perpendicular to v are
aŒ = a cos 21° = 10.58 m>s22cos 21° = 0.54 m>s2
S
a
21°
Perpendicular
component of acceleration
aŒ
Parallel component of acceleration
Position of rover at t = 2.0 s
a#
Path of rover
eVaLuate: The parallel component aΠis positive (in the same
S
direction as v), which means that the speed is increasing at this
instant. The value aΠ= + 0.54 m>s2 tells us that the speed is
increasing at this instant at a rate of 0.54 m>s per second. The
perpendicular component a# is not zero, which means that at this
instant the rover is turning—that is, it is changing direction and
following a curved path.
a# = a sin 21° = 10.58 m>s22sin 21° = 0.21 m>s2
Solution
ConCEptual ExamplE 3.4
aCCElEration of a skiEr
A skier moves along a ski-jump ramp (Fig. 3.14a). The ramp is
straight from point A to point C and curved from point C onward.
The skier speeds up as she moves downhill from point A to
point E, where her speed is maximum. She slows down after passing point E. Draw the direction of the acceleration vector at each
of the points B, D, E, and F.
3.14 (a) The skier’s path. (b) Our solution.
(a)
Direction
of motion
B
soLution
Figure 3.14b shows our solution. At point B the skier is moving
in a straight line with increasing speed, so her acceleration points
downhill, in the same direction as her velocity. At points D, E, and F
the skier is moving along a curved path, so her acceleration has
a component perpendicular to the path (toward the concave side
of the path) at each of these points. At point D there is also an
acceleration component in the direction of her motion because
she is speeding up. So the acceleration vector points ahead of the
normal to her path at point D. At point E, the skier’s speed is
instantaneously not changing; her speed is maximum at this point,
so its derivative is zero. There is therefore no parallel component
S
of a, and the acceleration is perpendicular to her motion. At point F
there is an acceleration component opposite to the direction of
her motion because she’s slowing down. The acceleration vector
therefore points behind the normal to her path.
In the next section we’ll consider the skier’s acceleration after
she flies off the ramp.
A
(b)
C
D
E
F
3.3 Projectile Motion
test your understanding of section 3.2 A sled travels over the crest of
a snow-covered hill. The sled slows down as it climbs up one side of the hill and gains
speed as it descends on the other side. Which of the vectors (1 through 9) in the figure
correctly shows the direction of the sled’s acceleration at the crest? (Choice 9 is that the
acceleration is zero.) ❙
75
3
2
4
1
5
Sled’s path
8
6
7
or 9: acceleration = 0
3.3 ProjectiLe Motion
A projectile is any body that is given an initial velocity and then follows a path
determined entirely by the effects of gravitational acceleration and air resistance.
A batted baseball, a thrown football, and a bullet shot from a rifle are all projectiles. The path followed by a projectile is called its trajectory.
To analyze the motion of a projectile, we’ll use an idealized model. We’ll
represent the projectile as a particle with an acceleration (due to gravity) that
is constant in both magnitude and direction. We’ll ignore the effects of air
resistance and the curvature and rotation of the earth. This model has limitations, however: We have to consider the earth’s curvature when we study the
flight of long-range missiles, and air resistance is of crucial importance to a
sky diver. Nevertheless, we can learn a lot from analysis of this simple model.
For the remainder of this chapter the phrase “projectile motion” will imply
that we’re ignoring air resistance. In Chapter 5 we’ll see what happens when
air resistance cannot be ignored.
Projectile motion is always confined to a vertical plane determined by the
direction of the initial velocity (Fig. 3.15). This is because the acceleration due
to gravity is purely vertical; gravity can’t accelerate the projectile sideways. Thus
projectile motion is two-dimensional. We will call the plane of motion the
xy-coordinate plane, with the x-axis horizontal and the y-axis vertically upward.
The key to analyzing projectile motion is that we can treat the x- and
y-coordinates separately. Figure 3.16 illustrates this for two projectiles: a red
ball dropped from rest and a yellow ball projected horizontally from the same
height. The figure shows that the horizontal motion of the yellow projectile
has no effect on its vertical motion. For both projectiles, the x-component of
acceleration is zero and the y-component is constant and equal to -g. (By
definition, g is always positive; with our choice of coordinate directions,
ay is negative.) So we can analyze projectile motion as a combination of
horizontal motion with constant velocity and vertical motion with constant
acceleration.
We can then express all the vector relationships for the projectile’s position,
velocity, and acceleration by separate equations for the horizontal and vertical
S
components. The components of a are
ax = 0
ay = -g
(projectile motion, no air resistance)
(3.13)
Since both the x-acceleration and y-acceleration are constant, we can use Eqs. (2.8),
(2.12), (2.13), and (2.14) directly. Suppose that at time t = 0 our particle is at the
point 1x0, y02 and its initial velocity at this time has components v0x and v0y . The
components of acceleration are ax = 0, ay = -g. Considering the x-motion first,
we substitute 0 for ax in Eqs. (2.8) and (2.12). We find
vx = v0x
(3.14)
x = x0 + v0x t
(3.15)
For the y-motion we substitute y for x, vy for vx , v0y for v0x , and ay = -g for ax :
vy = v0y - gt
(3.16)
y = y0 + v0y t - 12 gt 2
(3.17)
Demo
3.15 The trajectory of an idealized
projectile.
• A projectile moves in a vertical plane that
S
contains the initial velocity vector v0.
S
• Its trajectory depends only on v0 and
on the downward acceleration due to gravity.
y
Trajectory
S
v0
S
a
ax = 0, ay = -g
O
3.16 The red ball is dropped from rest,
and the yellow ball is simultaneously
projected horizontally.
The images
of the balls
are recorded
at equal
time intervals.
y
x
• At any time the two balls have different
x-coordinates and x-velocities but the same
y-coordinate, y-velocity, and y-acceleration.
• The horizontal motion of the yellow ball has
no effect on its vertical motion.
x
76
ChApTer 3 Motion in Two or Three Dimensions
3.17 If air resistance is negligible, the trajectory of a projectile is a combination of horizontal motion with
constant velocity and vertical motion with constant acceleration.
At the top of the trajectory, the projectile has zero vertical
velocity (vy = 0), but its vertical acceleration is still -g.
S
v2
y
S
v1
v1y
v1y
v3x
a
v1x
a
v3y
S
v3
ay = -g
S
v0
v0y
v0y
a0
O
Vertically, the projectile
v3y is in constant-acceleration
motion in response to the
earth’s gravitational pull.
Thus its vertical velocity
changes by equal amounts
during equal time intervals.
x
v0x
v0x
v1x
v2x
v3x
Horizontally, the projectile is in constant-velocity motion: Its horizontal acceleration
is zero, so it moves equal x-distances in equal time intervals.
3.18 The initial velocity components v0x
and v0y of a projectile (such as a kicked
soccer ball) are related to the initial speed
v0 and initial angle a0 .
y
S
v0
x
O
y
S
v0
It’s usually simplest to take the initial position 1at t = 02 as the origin; then
x0 = y0 = 0. This might be the position of a ball at the instant it leaves the hand
of the person who throws it or the position of a bullet at the instant it leaves the
gun barrel.
Figure 3.17 shows the trajectory of a projectile that starts at (or passes
through) the origin at time t = 0, along with its position, velocity, and velocity
components at equal time intervals. The x-velocity vx is constant; the y-velocity vy
changes by equal amounts in equal times, just as if the projectile were launched
vertically with the same initial y-velocity.
S
We can also represent the initial velocity v0 by its magnitude v0 (the initial
speed) and its angle a0 with the positive x-axis (Fig. 3.18). In terms of these
quantities, the components v0x and v0y of the initial velocity are
v0y = v0 sin a0
a0
v0x = v0 cos a0
x
v0x = v0 cos a0
v0y = v0 sin a0
(3.18)
If we substitute Eqs. (3.18) into Eqs. (3.14) through (3.17) and set x0 = y0 = 0,
we get the following equations. They describe the position and velocity of the
projectile in Fig. 3.17 at any time t:
Demo
Demo
PhET: Projectile Motion
Demo
Coordinates at time t of
a projectile (positive
y-direction is upward,
and x = y = 0 at t = 0)
x = 1v0 cos a02t
Speed
at t = 0
Direction
at t = 0
Time
y = 1v0 sin a02t - 12 gt 2
vx = v0 cos a0
Velocity components at
time t of a projectile
(positive y-direction
is upward)
(3.19)
Speed
at t = 0
Direction
at t = 0
vy = v0 sin a0 - gt
(3.20)
Acceleration
due to gravity:
Note g 7 0.
Time
(3.21)
(3.22)
3.3 projectile Motion
77
We can get a lot of information from Eqs. (3.19) through (3.22). For example,
the distance r from the origin to the projectile at any time t is
r = 2x 2 + y2
(3.23)
The projectile’s speed (the magnitude of its velocity) at any time is
v =
2v x2
v y2
+
(3.24)
The direction of the velocity, in terms of the angle a it makes with the positive
x-direction (see Fig. 3.17), is
tan a =
vy
Demo
3.19 The nearly parabolic trajectories of a
bouncing ball.
Successive images of the ball are
separated by equal time intervals.
Successive peaks decrease
in height because the ball
loses energy with
each bounce.
(3.25)
vx
S
The velocity vector v is tangent to the trajectory at each point.
We can derive an equation for the trajectory’s shape in terms of x and y by
eliminating t. From Eqs. (3.19) and (3.20), we find t = x>1v0 cos a02 and
y = 1tan a02x -
g
2v 02 cos2 a0
x2
(3.26)
Don’t worry about the details of this equation; the important point is its general
form. Since v0, tan a0, cos a0, and g are constants, Eq. (3.26) has the form
y = bx - cx 2
where b and c are constants. This is the equation of a parabola. In our simple
model of projectile motion, the trajectory is always a parabola (Fig. 3.19).
When air resistance isn’t negligible and has to be included, calculating the trajectory becomes a lot more complicated; the effects of air resistance depend on
velocity, so the acceleration is no longer constant. Figure 3.20 shows a computer
simulation of the trajectory of a baseball both without air resistance and with
air resistance proportional to the square of the baseball’s speed. We see that air
resistance has a very large effect; the projectile does not travel as far or as high,
and the trajectory is no longer a parabola.
soLution
Figure 3.21b shows our answer. The skier’s acceleration changed
from point to point while she was on the ramp. But as soon as she
G
F
H
Baseball’s initial velocity:
v0 = 50 m>s, a0 = 53.1°
50
O
-50
-100
100
200
With air
resistance
300
x (m)
No air
resistance
leaves the ramp, she becomes a projectile. So at points G, H, and I,
and indeed at all points after she leaves the ramp, the skier’s
acceleration points vertically downward and has magnitude g.
No matter how complicated the acceleration of a particle before
it becomes a projectile, its acceleration as a projectile is given by
ax = 0, ay = -g.
3.21 (a) The skier’s path during the jump. (b) Our solution.
(a)
y (m)
100
Solution
ConCEptual ExamplE 3.5 aCCElEration of a skiEr, ContinuEd
Let’s consider again the skier in Conceptual Example 3.4. What
is her acceleration at each of the points G, H, and I in Fig. 3.21a
after she flies off the ramp? Neglect air resistance.
3.20 Air resistance has a large cumulative
effect on the motion of a baseball. In this
simulation we allow the baseball to fall
below the height from which it was thrown
(for example, the baseball could have been
thrown from a cliff).
I
(b)
78
Chapter 3 Motion in two or three Dimensions
problEm-solving stratEgy 3.1
projECtilE motion
note: The strategies we used in Sections 2.4 and 2.5 for straight-
line, constant-acceleration problems are also useful here.
identify the relevant concepts: The key concept is that throughout projectile motion, the acceleration is downward and has a
constant magnitude g. Projectile-motion equations don’t apply to
throwing a ball, because during the throw the ball is acted on by
both the thrower’s hand and gravity. These equations apply only
after the ball leaves the thrower’s hand.
set uP the problem using the following steps:
1. Define your coordinate system and make a sketch showing
your axes. It’s almost always best to make the x-axis horizontal
and the y-axis vertical, and to choose the origin to be where
the body first becomes a projectile (for example, where a ball
leaves the thrower’s hand). Then the components of acceleration are ax = 0 and ay = - g, as in Eq. (3.13); the initial position is x0 = y0 = 0; and you can use Eqs. (3.19) through (3.22).
(If you choose a different origin or axes, you’ll have to modify
these equations.)
2. List the unknown and known quantities, and decide which
unknowns are your target variables. For example, you might
be given the initial velocity (either the components or the
magnitude and direction) and asked to find the coordinates
and velocity components at some later time. Make sure that
execute the solution: Find the target variables using the equa-
tions you chose. Resist the temptation to break the trajectory into
segments and analyze each segment separately. You don’t have to
start all over when the projectile reaches its highest point! It’s
almost always easier to use the same axes and time scale throughout the problem. If you need numerical values, use g = 9.80 m>s2.
Remember that g is positive!
eVaLuate your answer: Do your results make sense? Do the
numerical values seem reasonable?
Solution
ExamplE 3.6
you have as many equations as there are target variables to be
found. In addition to Eqs. (3.19) through (3.22), Eqs. (3.23)
through (3.26) may be useful.
3. State the problem in words and then translate those words into
symbols. For example, when does the particle arrive at a certain
point? (That is, at what value of t?) Where is the particle when
its velocity has a certain value? (That is, what are the values of
x and y when vx or vy has the specified value?) Since vy = 0
at the highest point in a trajectory, the question “When does
the projectile reach its highest point?” translates into “What is
the value of t when vy = 0?” Similarly, “When does the projectile return to its initial elevation?” translates into “What is
the value of t when y = y 0?”
a body projECtEd horizontally
A motorcycle stunt rider rides off the edge of a cliff. Just at the
edge his velocity is horizontal, with magnitude 9.0 m>s. Find the
motorcycle’s position, distance from the edge of the cliff, and
velocity 0.50 s after it leaves the edge of the cliff.
at t = 0.50 s, we use Eqs. (3.19) and (3.20); we then find the distance from the origin using Eq. (3.23). Finally, we use Eqs. (3.21)
and (3.22) to find the velocity components at t = 0.50 s.
execute: From Eqs. (3.19) and (3.20), the motorcycle’s x- and
y-coordinates at t = 0.50 s are
soLution
x = v0x t = 19.0 m>s210.50 s2 = 4.5 m
identify and set uP: Figure 3.22 shows our sketch of the tra-
jectory of motorcycle and rider. He is in projectile motion as soon
as he leaves the edge of the cliff, which we take to be the origin (so
S
x0 = y0 = 0). His initial velocity v0 at the edge of the cliff is horizontal (that is, a0 = 0), so its components are v0x = v0 cos a0 =
9.0 m>s and v0y = v0 sin a0 = 0. To find the motorcycle’s position
y = - 12 gt 2 = - 12 19.80 m>s2210.50 s22 = - 1.2 m
The negative value of y shows that the motorcycle is below its
starting point.
From Eq. (3.23), the motorcycle’s distance from the origin at
t = 0.50 s is
3.22 Our sketch for this problem.
At this point, the bike and
rider become a projectile.
are
r = 2x 2 + y2 = 214.5 m22 + 1- 1.2 m22 = 4.7 m
From Eqs. (3.21) and (3.22), the velocity components at t = 0.50 s
vx = v0x = 9.0 m>s
vy = - gt = 1-9.80 m>s2210.50 s2 = -4.9 m>s
The motorcycle has the same horizontal velocity vx as when it left
the cliff at t = 0, but in addition there is a downward (negative)
vertical velocity vy . The velocity vector at t = 0.50 s is
S
v = vxnd + vyne = 19.0 m>s2dn + 1-4.9 m>s2en
3.3 projectile Motion
From Eqs. (3.24) and (3.25), at t = 0.50 s the velocity has
magnitude v and angle a given by
v = 2v x2 + v y2 = 219.0 m>s22 + 1-4.9 m>s22 = 10.2 m>s
a = arctan
vy
vx
= arctan a
- 4.9 m>s
b = -29°
9.0 m >s
79
eVaLuate: Just as in Fig. 3.17, the motorcycle’s horizontal motion is unchanged by gravity; the motorcycle continues to move
horizontally at 9.0 m>s, covering 4.5 m in 0.50 s. The motorcycle
initially has zero vertical velocity, so it falls vertically just like a
body released from rest and descends a distance 12 gt 2 = 1.2 m in
0.50 s.
The motorcycle is moving at 10.2 m>s in a direction 29° below the
horizontal.
HEigHt And rAngE of A projEctilE i: A bAttEd bAsEbAll
A batter hits a baseball so that it leaves the bat at speed
v0 = 37.0 m>s at an angle a0 = 53.1°. (a) Find the position of
the ball and its velocity (magnitude and direction) at t = 2.00 s.
(b) Find the time when the ball reaches the highest point of its
flight, and its height h at this time. (c) Find the horizontal range
R—that is, the horizontal distance from the starting point to where
the ball hits the ground—and the ball’s velocity just before it hits.
soLution
identify and set uP: As Fig. 3.20 shows, air resistance strongly
affects the motion of a baseball. For simplicity, however, we’ll ignore air resistance here and use the projectile-motion equations to
describe the motion. The ball leaves the bat at t = 0 a meter or so
above ground level, but we’ll ignore this distance and assume that
it starts at ground level 1y0 = 02. Figure 3.23 shows our sketch
of the ball’s trajectory. We’ll use the same coordinate system as
in Figs. 3.17 and 3.18, so we can use Eqs. (3.19) through (3.22).
Our target variables are (a) the position and velocity of the ball
2.00 s after it leaves the bat, (b) the time t when the ball is at its
maximum height (that is, when vy = 0) and the y-coordinate at
this time, and (c) the x-coordinate when the ball returns to ground
level 1y = 02 and the ball’s vertical component of velocity then.
execute: (a) We want to find x, y, vx , and vy at t = 2.00 s. The
initial velocity of the ball has components
v0x = v0 cos a0 = 137.0 m>s2cos 53.1° = 22.2 m>s
v0y = v0 sin a0 = 137.0 m>s2sin 53.1° = 29.6 m>s
vx = v0x = 22.2 m>s
vy = v0y - gt = 29.6 m>s - 19.80 m>s2212.00 s2 = 10.0 m>s
The y-component of velocity is positive at t = 2.00 s, so the ball
is still moving upward (Fig. 3.23). From Eqs. (3.24) and (3.25), the
magnitude and direction of the velocity are
v = 2v x2 + v y2 = 2122.2 m>s22 + 110.0 m>s22 = 24.4 m>s
a = arctan a
x = v0x t = 122.2 m>s212.00 s2 = 44.4 m
vy = v0y - gt1 = 0
t1 =
= 129.6 m>s212.00 s2 -
3.23 Our sketch for this problem.
2
v0y
g
=
29.6 m>s
9.80 m>s2
= 3.02 s
The height h at the highest point is the value of y at time t1:
h = v0y t1 - 12 gt 12
= 129.6 m>s213.02 s2 - 12 19.80 m>s2213.02 s22 = 44.7 m
(c) We’ll find the horizontal range in two steps. First, we find
the time t2 when y = 0 (the ball is at ground level):
y = 0 = v0y t2 - 12 gt 22 = t2 1 v0y - 12 gt2 2
This is a quadratic equation for t2. It has two roots:
1 2
2 gt
1
2 19.80
10.0 m>s
b = arctan 0.450 = 24.2°
22.2 m>s
The ball is moving at 24.4 m/s in a direction 24.2° above the
horizontal.
(b) At the highest point, the vertical velocity vy is zero. Call the
time when this happens t1; then
From Eqs. (3.19) through (3.22),
y = v0y t -
Solution
ExAmplE 3.7
2
m>s 212.00 s2 = 39.6 m
t2 = 0
and
t2 =
2v0y
g
=
2129.6 m>s2
9.80 m>s2
= 6.04 s
The ball is at y = 0 at both times. The ball leaves the ground at
t2 = 0, and it hits the ground at t2 = 2v0y >g = 6.04 s.
The horizontal range R is the value of x when the ball returns
to the ground at t2 = 6.04 s:
R = v0x t2 = 122.2 m>s216.04 s2 = 134 m
The vertical component of velocity when the ball hits the
ground is
vy = v0y - gt2 = 29.6 m>s - 19.80 m>s2216.04 s2
= - 29.6 m>s
Continued
80
Chapter 3 Motion in two or three Dimensions
That is, vy has the same magnitude as the initial vertical velocity v0y
but the opposite direction (down). Since vx is constant, the angle
a = -53.1° (below the horizontal) at this point is the negative of
the initial angle a0 = 53.1°.
eVaLuate: It’s often useful to check results by getting them in a
different way. For example, we can also find the maximum height
in part (b) by applying the constant-acceleration formula Eq. (2.13)
to the y-motion:
v y2 = v 0y 2 + 2ay 1y - y02 = v 0y 2 - 2g1y - y02
At the highest point, vy = 0 and y = h. Solve this equation for h;
you should get the answer that we obtained in part (b). (Do you?)
hEight and rangE of a projECtilE ii: maximum hEight, maximum rangE
Find the maximum height h and horizontal range R (see Fig. 3.23)
of a projectile launched with speed v0 at an initial angle a0 between
0 and 90°. For a given v0, what value of a0 gives maximum height?
What value gives maximum horizontal range?
soLution
identify and set uP: This is almost the same as parts (b) and (c)
of Example 3.7, except that now we want general expressions for h
and R. We also want the values of a0 that give the maximum values of h and R. In part (b) of Example 3.7 we found that the projectile reaches the high point of its trajectory (so that vy = 0) at time
t1 = v0y >g, and in part (c) we found that the projectile returns to
its starting height (so that y = y0) at time t2 = 2v0y >g = 2t1. We’ll
use Eq. (3.20) to find the y-coordinate h at t1 and Eq. (3.19) to find
the x-coordinate R at time t2. We’ll express our answers in terms
of the launch speed v0 and launch angle a0 by using Eqs. (3.18).
execute: From Eqs. (3.18), v0x = v0 cos a0 and v0y = v0 sin a0.
Hence we can write the time t1 when vy = 0 as
t1 =
v0y
g
=
v0 sin a0
g
Equation (3.20) gives the height y = h at this time:
h = 1v0 sin a02a
v0 sin a0
v0 sin a0 2
v 02 sin2 a0
b - 12 ga
b =
g
g
2g
For a given launch speed v0, the maximum value of h occurs for
sin a0 = 1 and a0 = 90°—that is, when the projectile is launched
straight up. (If it is launched horizontally, as in Example 3.6,
a0 = 0 and the maximum height is zero!)
The time t2 when the projectile hits the ground is
t2 =
2v0y
g
=
2v0 sin a0
g
The horizontal range R is the value of x at this time. From Eq. (3.19),
this is
R = 1v0 cos a02t2 = 1v0 cos a02
2v0 sin a0
v 02 sin 2a0
=
g
g
Solution
ExamplE 3.8
Note that the time to hit the ground, t2 = 6.04 s , is exactly
twice the time to reach the highest point, t1 = 3.02 s. Hence the
time of descent equals the time of ascent. This is always true if
the starting point and endpoint are at the same elevation and if air
resistance can be ignored.
Note also that h = 44.7 m in part (b) is comparable to the
61.0-m height above second base of the roof at Marlins Park in
Miami, and the horizontal range R = 134 m in part (c) is greater
than the 99.7-m distance from home plate to the right-field fence
at Safeco Field in Seattle. In reality, due to air resistance (which
we have ignored) a batted ball with the initial speed and angle
we’ve used here won’t go as high or as far as we’ve calculated (see
Fig. 3.20).
(We used the trigonometric identity 2 sin a0 cos a0 = sin 2a0,
found in Appendix B.) The maximum value of sin 2a0 is 1; this
occurs when 2a0 = 90°, or a0 = 45°. This angle gives the maximum range for a given initial speed if air resistance can be ignored.
eVaLuate: Figure 3.24 is based on a composite photograph of
three trajectories of a ball projected from a small spring gun at
angles of 30°, 45°, and 60°. The initial speed v0 is approximately
the same in all three cases. The horizontal range is greatest for
the 45° angle. The ranges are nearly the same for the 30° and
60° angles: Can you prove that for a given value of v0 the range is
the same for both an initial angle a0 and an initial angle 90° - a0 ?
(This is not the case in Fig. 3.24 due to air resistance.)
caution height and range of a projectile We don’t recommend
memorizing the above expressions for h and R. They are applicable
only in the special circumstances we’ve described. In particular,
you can use the expression for the range R only when launch and
landing heights are equal. There are many end-of-chapter problems to which these equations do not apply. ❙
3.24 A launch angle of 45° gives the maximum horizontal range.
The range is shorter with launch angles of 30° and 60°.
A 45° launch angle gives the greatest range;
other angles fall shorter.
Launch
angle:
a0 = 30°
a0 = 45°
a0 = 60°
3.3 Projectile Motion
Solution
ExamplE 3.9
81
diffErEnt initial and final HEigHts
You throw a ball from your window 8.0 m above the ground.
When the ball leaves your hand, it is moving at 10.0 m>s at an
angle of 20° below the horizontal. How far horizontally from your
window will the ball hit the ground? Ignore air resistance.
execute: To determine t, we rewrite Eq. (3.20) in the standard
form for a quadratic equation for t:
1 2
2 gt
- 1v0 sin a02t + y = 0
The roots of this equation are
soLution
identify and set uP: As in Examples 3.7 and 3.8, we want to
find the horizontal coordinate of a projectile when it is at a given
y-value. The difference here is that this value of y is not the same
as the initial value. We again choose the x-axis to be horizontal
and the y-axis to be upward, and place the origin of coordinates
at the point where the ball leaves your hand (Fig. 3.25). We have
v0 = 10.0 m>s and a0 = - 20° (the angle is negative because the
initial velocity is below the horizontal). Our target variable is the
value of x when the ball reaches the ground at y = -8.0 m. We’ll
use Eq. (3.20) to find the time t when this happens and then use
Eq. (3.19) to find the value of x at this time.
3.25 Our sketch for this problem.
t =
=
v0 sin a0 { 31-v0 sin a022 - 4 1 12 g 2 y
2 1 12 g 2
v0 sin a0 { 2v 02 sin2 a0 - 2gy
g
J
=
110.0 m>s2 sin1-20°2
R
{ 2110.0 m>s22 sin2 1-20°2 - 219.80 m>s221- 8.0 m2
= -1.7 s
9.80 m>s2
or
0.98 s
We discard the negative root, since it refers to a time before the
ball left your hand. The positive root tells us that the ball reaches
the ground at t = 0.98 s. From Eq. (3.19), the ball’s x-coordinate
at that time is
x = 1v0 cos a02t = 110.0 m>s23cos1-20°2410.98 s2 = 9.2 m
The ball hits the ground a horizontal distance of 9.2 m from your
window.
eVaLuate: The root t = -1.7 s is an example of a “fictional” so-
lution to a quadratic equation. We discussed these in Example 2.8
in Section 2.5; review that discussion.
Solution
ExamplE 3.10
tHE zookEEpEr and tHE monkEy
A monkey escapes from the zoo and climbs a tree. After failing
to entice the monkey down, the zookeeper fires a tranquilizer
dart directly at the monkey (Fig. 3.26). The monkey lets go at the
instant the dart leaves the gun. Show that the dart will always hit
the monkey, provided that the dart reaches the monkey before he
hits the ground and runs away.
soLution
identify and set uP: We have two bodies in projectile motion:
the dart and the monkey. They have different initial positions and
initial velocities, but they go into projectile motion at the same
time t = 0. We’ll first use Eq. (3.19) to find an expression for the
time t when the x-coordinates xmonkey and xdart are equal. Then
we’ll use that expression in Eq. (3.20) to see whether ymonkey
and ydart are also equal at this time; if they are, the dart hits the
monkey. We make the usual choice for the x- and y-directions, and
place the origin of coordinates at the muzzle of the tranquilizer
gun (Fig. 3.26).
execute: The monkey drops straight down, so xmonkey = d at all
times. From Eq. (3.19), xdart = 1v0 cos a02t. We solve for the time t
when these x-coordinates are equal:
d = 1v0 cos a02t
so
t =
d
v0 cos a0
We must now show that ymonkey = ydart at this time. The monkey
is in one-dimensional free fall; its position at any time is given by
Eq. (2.12), with appropriate symbol changes. Figure 3.26 shows
that the monkey’s initial height above the dart-gun’s muzzle is
ymonkey - 0 = d tan a0, so
ymonkey = d tan a0 - 12 gt 2
From Eq. (3.20),
ydart = 1v0 sin a02t - 12 gt 2
Continued
82
ChAPTer 3 Motion in Two or Three Dimensions
3.26 The tranquilizer dart hits the falling monkey.
Dashed arrows show how far the dart and monkey have fallen at
specific times relative to where they would be without gravity.
At any time, they have fallen by the same amount.
y
Without gravity
• The monkey remains in its initial position.
• The dart travels straight to the monkey.
• Therefore, the dart hits the monkey.
Trajectory of dart
without gravity
Monkey’s
fall
Dart’s
fall
Dart’s
d tan a0 fall
Dart’s fall
v0
a0
Trajectory of dart
with gravity
O
x
d
With gravity
• The monkey falls straight down.
• At any time t, the dart has fallen by the same amount
as the monkey relative to where either would be in the
absence of gravity: ∆ydart = ∆ymonkey = - 12 gt 2.
• Therefore, the dart always hits the monkey.
Comparing these two equations, we see that we’ll have
ymonkey = ydart (and a hit) if d tan a0 = 1v0 sin a02t when the two
x-coordinates are equal. To show that this happens, we replace t
with d>1v0 cos a02, the time when xmonkey = xdart. Sure enough,
1v0 sin a02t = 1v0 sin a02
d
= d tan a0
v0 cos a0
equal; a dart aimed at the monkey always hits it, no matter what v0
is (provided the monkey doesn’t hit the ground first). This result is
independent of the value of g, the acceleration due to gravity. With
no gravity 1g = 02, the monkey would remain motionless, and the
dart would travel in a straight line to hit him. With gravity, both
fall the same distance gt 2>2 below their t = 0 positions, and the
dart still hits the monkey (Fig. 3.26).
eVaLuate: We’ve proved that the y-coordinates of the dart and
the monkey are equal at the same time that their x-coordinates are
A
P
B
C
test your understanding of section 3.3 In Example 3.10, suppose the
tranquilizer dart has a relatively low muzzle velocity so that the dart reaches a maximum
height at a point P before striking the monkey, as shown in the figure. When the dart is
at point P, will the monkey be (i) at point A (higher than P), (ii) at point B (at the same
height as P), or (iii) at point C (lower than P)? Ignore air resistance. ❙
3.4 Motion in a circLe
When a particle moves along a curved path, the direction of its velocity changes.
As we saw in Section 3.2, this means that the particle must have a component of
acceleration perpendicular to the path, even if its speed is constant (see Fig. 3.11b).
In this section we’ll calculate the acceleration for the important special case of
motion in a circle.
uniform circular Motion
When a particle moves in a circle with constant speed, the motion is called
uniform circular motion. A car rounding a curve with constant radius at constant
speed, a satellite moving in a circular orbit, and an ice skater skating in a circle
83
3.4 Motion in a Circle
3.27 A car moving along a circular path. If the car is in uniform circular motion as in (a), the speed is constant
and the acceleration is directed toward the center of the circular path (compare Fig. 3.12).
(a) Uniform circular motion: Constant speed
along a circular path
(b) Car speeding up along a circular path
Component of acceleration parallel to velocity:
Changes car’s speed
(c) Car slowing down along a circular path
S
S
v
S
v
Acceleration is exactly
S perpendicular to velocity;
a
no parallel component
v
Component of acceleration
perpendicular to velocity:
Changes car’s direction
S
a
S
a
Component of acceleration perpendicular to
velocity: Changes car’s direction
To center of circle
with constant speed are all examples of uniform circular motion (Fig. 3.27a;
compare Fig. 3.12a). There is no component of acceleration parallel (tangent) to
the path; otherwise, the speed would change. The acceleration vector is perpendicular (normal) to the path and hence directed inward (never outward!) toward
the center of the circular path. This causes the direction of the velocity to change
without changing the speed.
We can find a simple expression for the magnitude of the acceleration in uniform circular motion. We begin with Fig. 3.28a, which shows a particle moving
with constant speed in a circular path of radius R with center at O. The particle
moves a distance ∆s from P1 to P2 in a time interval ∆t. Figure 3.28b shows the
S
vector change in velocity ∆v during this interval.
S
The angles labeled ∆f in Figs. 3.28a and 3.28b are the same because v1 is
S
perpendicular to the line OP1 and v2 is perpendicular to the line OP2 . Hence the
triangles in Figs. 3.28a and 3.28b are similar. The ratios of corresponding sides
of similar triangles are equal, so
0 ∆vS 0
v1
=
∆s
R
0 ∆vS 0 =
or
Component of acceleration parallel
to velocity: Changes car’s speed
S
3.28 Finding the velocity change ∆v,
S
average acceleration aav, and instantaneous
S
acceleration arad for a particle moving in a
circle with constant speed.
(a) A particle moves a distance ∆s at
constant speed along a circular path.
S
v2
S
v1
P2
P1
v1
∆s
R
∆s
R
∆f
R
The magnitude aav of the average acceleration during ∆t is therefore
0 ∆v 0
S
aav =
∆t
=
O
v1 ∆s
R ∆t
S
The magnitude a of the instantaneous acceleration a at point P1 is the limit of
this expression as we take point P2 closer and closer to point P1:
a = lim
∆t S 0
(b) The corresponding change in velocity and
average acceleration
v1 ∆s
v1
∆s
=
lim
R ∆t
R ∆t S 0 ∆t
∆f
If the time interval ∆t is short, ∆s is the distance the particle moves along its
curved path. So the limit of ∆s>∆t is the speed v1 at point P1 . Also, P1 can be any
point on the path, so we can drop the subscript and let v represent the speed at
any point. Then
These two triangles
are similar.
S
S
v1
∆v
S
v2
O
(c) The instantaneous acceleration
S
v
Magnitude of acceleration
of an object in
uniform circular motion
arad =
v2
R
Speed of object
Radius of object’s
circular path
(3.27)
S
The subscript “rad” is a reminder that the direction of the instantaneous acceleration at each point is always along a radius of the circle (toward the center of the
circle; see Figs. 3.27a and 3.28c). So in uniform circular motion, the magnitude
arad
The instantaneous acceleration
in uniform circular motion
always points toward the
R
center of the circle.
O
84
ChAPTer 3 Motion in Two or Three Dimensions
3.29 Acceleration and velocity (a) for a
particle in uniform circular motion and
(b) for a projectile with no air resistance.
(a) Uniform circular motion
S
v
S
v
S
arad
S
v
S
arad
S
S
v
arad
S
arad
S
arad
S
v
acceleration on a curved Path
When students were given a problem
about an object following a curved path
(not necessarily the parabolic path of a
projectile), more than 46% gave an
incorrect answer. Common errors:
●
●
Confusion between the acceleration
S
S
vector a and the velocity vector v.
S
Remember that a is the rate of change of
S
S
S
v, and on a curved path a and v cannot
be in the same direction (see Fig. 3.12).
S
Confusion about the direction of a. If
S
the path is curved, a always has a
component toward the inside of the
curve (see Fig. 3.12).
Velocity and acceleration are perpendicular
only at the peak of the trajectory.
S
S
v
v
arad
S
data SpeakS
(b) Projectile motion
Acceleration has
constant magnitude but varying
direction.
S
v
S
v
S
a
S
v
Velocity and
acceleration
are always
perpendicular.
S
S
a
a
Acceleration is
constant in magnitude
and direction.
S
a
S
v
S
a
arad of the instantaneous acceleration is equal to the square of the speed v
S
divided by the radius R of the circle. Its direction is perpendicular to v and
inward along the radius (Fig. 3.29a).
Because the acceleration in uniform circular motion is always directed toward
the center of the circle, it is sometimes called centripetal acceleration. The word
“centripetal” is derived from two Greek words meaning “seeking the center.”
caution uniform circular motion vs. projectile motion Notice the differences between
acceleration in uniform circular motion (Fig. 3.29a) and acceleration in projectile motion (Fig. 3.29b). It’s true that in both kinds of motion the magnitude of acceleration is
S
the same at all times. However, in uniform circular motion the direction of a changes
continuously—it always points toward the center of the circle. In projectile motion, the
S
direction of a remains the same at all times. ❙
We can also express the magnitude of the acceleration in uniform circular
motion in terms of the period T of the motion, the time for one revolution (one
complete trip around the circle). In a time T the particle travels a distance equal
to the circumference 2pR of the circle, so its speed is
v =
2pR
T
(3.28)
When we substitute this into Eq. (3.27), we obtain the alternative expression
PhET: Ladybug Revolution
PhET: Motion in 2D
Magnitude of acceleration
of an object in
uniform circular motion
arad =
4p2R
T2
Radius of object’s circular path
Period of motion
Solution
ExamplE 3.11 CEntripEtal aCCElEration on a CurvEd road
An Aston Martin V8 Vantage sports car has a “lateral acceleration”
of 0.96g = 10.96219.8 m>s22 = 9.4 m>s2. This is the maximum
centripetal acceleration the car can sustain without skidding out
of a curved path. If the car is traveling at a constant 40 m>s (about
89 mi>h, or 144 km>h) on level ground, what is the radius R of the
tightest unbanked curve it can negotiate?
soLution
identify, set uP, and execute: The car is in uniform circular
motion because it’s moving at a constant speed along a curve that
is a segment of a circle. Hence we can use Eq. (3.27) to solve for
the target variable R in terms of the given centripetal accelera-
(3.29)
ion arad and speed v:
R =
140 m>s22
v2
=
= 170 m 1about 560 ft2
arad
9.4 m>s2
This is the minimum turning radius because arad is the maximum
centripetal acceleration.
eVaLuate: The minimum turning radius R is proportional to the
square of the speed, so even a small reduction in speed can make
R substantially smaller. For example, reducing v by 20% (from
40 m>s to 32 m>s) would decrease R by 36% (from 170 m to 109 m).
Another way to make the minimum turning radius smaller is to
bank the curve. We’ll investigate this option in Chapter 5.
3.4 Motion in a Circle
Solution
ExAmplE 3.12 cEntripEtAl AccElErAtion on A cArnivAl ridE
Passengers on a carnival ride move at constant speed in a horizontal circle of radius 5.0 m, making a complete circle in 4.0 s. What
is their acceleration?
eVaLuate: We can check this answer by using the second, roundabout approach. From Eq. (3.28), the speed is
v =
soLution
identify and set uP: The speed is constant, so this is uniform
circular motion. We are given the radius R = 5.0 m and the period T = 4.0 s, so we can use Eq. (3.29) to calculate the acceleration directly, or we can calculate the speed v by using Eq. (3.28)
and then find the acceleration by using Eq. (3.27).
execute: From Eq. (3.29),
arad =
4p215.0 m2
14.0 s22
= 12 m>s2 = 1.3g
arad =
d0v0
17.9 m>s22
v2
=
= 12 m>s2
R
5.0 m
S
As in Fig. 3.29a, the direction of a is always toward the center of
S
the circle. The magnitude of a is relatively mild as carnival rides
go; some roller coasters subject their passengers to accelerations
as great as 4g.
We have assumed throughout this section that the particle’s speed is constant
as it goes around the circle. If the speed varies, we call the motion nonuniform
circular motion. In nonuniform circular motion, Eq. (3.27) still gives the radial
component of acceleration arad = v 2>R, which is always perpendicular to the
instantaneous velocity and directed toward the center of the circle. But since the
speed v has different values at different points in the motion, the value of arad is
not constant. The radial (centripetal) acceleration is greatest at the point in the
circle where the speed is greatest.
In nonuniform circular motion there is also a component of acceleration that
is parallel to the instantaneous velocity (see Figs. 3.27b and 3.27c). This is the
component aΠthat we discussed in Section 3.2; here we call this component atan
to emphasize that it is tangent to the circle. The tangential component of acceleration atan is equal to the rate of change of speed. Thus
v2
R
2p15.0 m2
2pR
=
= 7.9 m>s
T
4.0 s
The centripetal acceleration is then
nonuniform circular Motion
arad =
85
application Watch Out: Tight
Curves Ahead! These roller coaster cars
are in nonuniform circular motion: They slow
down and speed up as they move around a
vertical loop. The large accelerations involved
in traveling at high speed around a tight loop
mean extra stress on the passengers’
circulatory systems, which is why people with
cardiac conditions are cautioned against
going on such rides.
S
and
atan =
dt
(nonuniform circular motion)
(3.30)
The tangential component is in the same direction as the velocity if the particle
is speeding up, and in the opposite direction if the particle is slowing down
(Fig. 3.30). If the particle’s speed is constant, atan = 0.
caution Uniform vs. nonuniform circular motion The two quantities
d0v0
S
dt
S
and
`
dv
`
dt
are not the same. The first, equal to the tangential acceleration, is the rate of change of
speed; it is zero whenever a particle moves with constant speed, even when its direction of
motion changes (such as in uniform circular motion). The second is the magnitude of the
vector acceleration; it is zero only when the particle’s acceleration vector is zero—that is,
when the particle moves in a straight line with constant speed. In uniform circular motion
0 dvS >dt 0 = arad = v 2>r; in nonuniform circular motion there is also a tangential compoS
nent of acceleration, so 0 dv>dt 0 = 2arad2 + atan2 . ❙
test your understanding of section 3.4 Suppose that the particle in
Fig. 3.30 experiences four times the acceleration at the bottom of the loop as it does at the
top of the loop. Compared to its speed at the top of the loop, is its speed at the bottom of
the loop (i) 12 times as great; (ii) 2 times as great; (iii) 2 12 times as great; (iv) 4 times
as great; or (v) 16 times as great? ❙
3.30 A particle moving in a vertical loop
with a varying speed, like a roller coaster car.
Speed slowest, arad minimum, atan zero
Speeding up; atan in
S
same direction as v
atan
S
v
Slowing down;
S
atan opposite to v
arad
S
a
0 aS 0 = arad
Speed fastest, arad maximum, atan zero
86
Chapter 3 Motion in two or three Dimensions
3.31 Airshow pilots face a complicated
problem involving relative velocities. They
must keep track of their motion relative to
the air (to maintain enough airflow over
the wings to sustain lift), relative to each
other (to keep a tight formation without
colliding), and relative to their audience
(to remain in sight of the spectators).
3.5 reLatiVe VeLocity
If you stand next to a one-way highway, all the cars appear to be moving forward.
But if you’re driving in the fast lane on that highway, slower cars appear to be
moving backward. In general, when two observers measure the velocity of the
same body, they get different results if one observer is moving relative to the
other. The velocity seen by a particular observer is called the velocity relative to
that observer, or simply relative velocity. In many situations relative velocity is
extremely important (Fig. 3.31).
We’ll first consider relative velocity along a straight line and then generalize
to relative velocity in a plane.
relative Velocity in one dimension
3.32 (a) A passenger walking in a train.
(b) The position of the passenger relative
to the cyclist’s frame of reference and the
train’s frame of reference.
(a)
P (passenger)
B (train)
A passenger walks with a velocity of 1.0 m>s along the aisle of a train that is
moving with a velocity of 3.0 m>s (Fig. 3.32a). What is the passenger’s velocity?
It’s a simple enough question, but it has no single answer. As seen by a second
passenger sitting in the train, she is moving at 1.0 m>s. A person on a bicycle
standing beside the train sees the walking passenger moving at 1.0 m>s +
3.0 m>s = 4.0 m>s. An observer in another train going in the opposite direction
would give still another answer. We have to specify which observer we mean,
and we speak of the velocity relative to a particular observer. The walking
passenger’s velocity relative to the train is 1.0 m>s, her velocity relative to the
cyclist is 4.0 m>s, and so on. Each observer, equipped in principle with a meter
stick and a stopwatch, forms what we call a frame of reference. Thus a frame of
reference is a coordinate system plus a time scale.
Let’s use the symbol A for the cyclist’s frame of reference (at rest with respect to the ground) and the symbol B for the frame of reference of the moving
train. In straight-line motion the position of a point P relative to frame A is given
by xP>A (the position of P with respect to A), and the position of P relative to
frame B is given by xP>B (Fig. 3.32b). The position of the origin of B with respect
to the origin of A is xB>A. Figure 3.32b shows that
xP>A = xP>B + xB>A
A (cyclist)
(b)
yA
yB
Cyclist's
frame
vB>A
Train’s
frame
Velocity of train
relative to cyclist
Position of passenger
in both frames
P
OA
OB
xB>A
xP>A
xP>B
xB ,
xA
(3.31)
In words, the coordinate of P relative to A equals the coordinate of P relative to B
plus the coordinate of B relative to A.
The x-velocity of P relative to frame A, denoted by vP>A@x, is the derivative of
xP>A with respect to time. We can find the other velocities in the same way. So the
time derivative of Eq. (3.31) gives us a relationship among the various velocities:
dxP>A
dt
Relative velocity
vP>A-x
along a line:
x-velocity of
P relative to A
=
dxP>B
dt
+
dxB>A
or
dt
= vP>B-x + vB>A-x
x-velocity of
P relative to B
(3.32)
x-velocity of
B relative to A
Getting back to the passenger on the train in Fig. 3.32a, we see that A is the
cyclist’s frame of reference, B is the frame of reference of the train, and point P
represents the passenger. Using the above notation, we have
vP>B@x = +1.0 m>s
vB>A@x = +3.0 m>s
From Eq. (3.32) the passenger’s velocity vP>A@x relative to the cyclist is
vP>A@x = +1.0 m>s + 3.0 m>s = +4.0 m>s
as we already knew.
3.5 relative Velocity
87
In this example, both velocities are toward the right, and we have taken
this as the positive x-direction. If the passenger walks toward the left relative
to the train, then vP>B@x = -1.0 m>s, and her x-velocity relative to the cyclist is
vP>A@x = -1.0 m>s + 3.0 m>s = +2.0 m>s. The sum in Eq. (3.32) is always an
algebraic sum, and any or all of the x-velocities may be negative.
When the passenger looks out the window, the stationary cyclist on the ground
appears to her to be moving backward; we call the cyclist’s velocity relative to
her vA>P@x . This is just the negative of the passenger’s velocity relative to the
cyclist, vP>A@x . In general, if A and B are any two points or frames of reference,
vA>B@x = -vB>A@x
problEm-solving strAtEgy 3.2
rElAtivE vElocity
identify the relevant concepts: Whenever you see the phrase
“velocity relative to” or “velocity with respect to,” it’s likely that
the concepts of relative velocity will be helpful.
set uP the problem: Sketch and label each frame of reference in
the problem. Each moving body has its own frame of reference;
in addition, you’ll almost always have to include the frame of
reference of the earth’s surface. (Statements such as “The car is
traveling north at 90 km>h” implicitly refer to the car’s velocity
relative to the surface of the earth.) Use the labels to help identify the target variable. For example, if you want to find the
x-velocity of a car 1C2 with respect to a bus 1B2, your target
variable is vC>B@x .
execute the solution: Solve for the target variable using Eq. (3.32).
(If the velocities aren’t along the same direction, you’ll need to use
the vector form of this equation, derived later in this section.) It’s
important to note the order of the double subscripts in Eq. (3.32):
vB>A@x means “x-velocity of B relative to A.” These subscripts obey
a kind of algebra. If we regard each one as a fraction, then the
fraction on the left side is the product of the fractions on the right
side: P>A = 1P>B21B>A2. You can apply this rule to any number
of frames of reference. For example, if there are three frames of
reference A, B, and C, Eq. (3.32) becomes
vP>A@x = vP>C@x + vC>B@x + vB>A@x
eVaLuate your answer: Be on the lookout for stray minus signs
in your answer. If the target variable is the x-velocity of a car relative to a bus 1vC>B@x2, make sure that you haven’t accidentally
calculated the x-velocity of the bus relative to the car 1vB>C@x2. If
you’ve made this mistake, you can recover by using Eq. (3.33).
Solution
ExAmplE 3.13
(3.33)
rElAtivE vElocity on A strAigHt roAd
You drive north on a straight two-lane road at a constant 88 km>h.
A truck in the other lane approaches you at a constant 104 km>h
(Fig. 3.33). Find (a) the truck’s velocity relative to you and (b) your
velocity relative to the truck. (c) How do the relative velocities
change after you and the truck pass each other? Treat this as a
one-dimensional problem.
3.33 Reference frames for you and the truck.
N
E
W
soLution
identify and set uP: In this problem about relative velocities along a line, there are three reference frames: you (Y), the
truck (T), and the earth’s surface (E). Let the positive x-direction
be north (Fig. 3.33). Then your x-velocity relative to the earth is
vY>E@x = + 88 km>h. The truck is initially approaching you, so
it is moving south and its x-velocity with respect to the earth is
vT>E@x = - 104 km>h. The target variables in parts (a) and (b) are
vT>Y@x and vY>T@x , respectively. We’ll use Eq. (3.32) to find the first
target variable and Eq. (3.33) to find the second.
S
execute: (a) To find vT>Y@x , we write Eq. (3.32) for the known
x
vT>E@x and rearrange:
Truck (T )
vT>E@x = vT>Y@x + vY>E@x
vT>Y@x = vT>E@x - vY>E@x
S
vY>E
Earth (E)
= -104 km>h - 88 km>h = - 192 km>h
S
vT>E
You (Y)
The truck is moving at 192 km>h in the negative x-direction
(south) relative to you.
Continued
88
Chapter 3 Motion in two or three Dimensions
the south relative to you, even though it is now moving away from
you instead of toward you.
(b) From Eq. (3.33),
vY>T@x = - vT>Y@x = - 1- 192 km>h2 = + 192 km>h
You are moving at 192 km>h in the positive x-direction (north)
relative to the truck.
(c) The relative velocities do not change after you and the truck
pass each other. The relative positions of the bodies don’t matter.
After it passes you the truck is still moving at 192 km>h toward
eVaLuate: To check your answer in part (b), use Eq. (3.32) directly
in the form vY>T@x = vY>E@x + vE>T@x . (The x-velocity of the earth
with respect to the truck is the opposite of the x-velocity of the
truck with respect to the earth: vE>T@x = - vT>E@x .) Do you get the
same result?
relative Velocity in two or three dimensions
Let’s extend the concept of relative velocity to include motion in a plane or in
space. Suppose that the passenger in Fig. 3.32a is walking not down the aisle of
the railroad car but from one side of the car to the other, with a speed of 1.0 m>s
(Fig. 3.34a). We can again describe the passenger’s position P in two frames of
reference: A for the stationary ground observer and B for the moving train. But
S
instead of coordinates x, we use position vectors r because the problem is now
two-dimensional. Then, as Fig. 3.34b shows,
S
S
S
rP>A = rP>B + rB>A
(3.34)
Just as we did before, we take the time derivative of this equation to get a relaS
tionship among the various velocities; the velocity of P relative to A is vP>A =
S
drP>A>dt and so on for the other velocities. We get
S
S
S
Relative velocity
vP>A = vP>B + vB>A
in space:
Velocity of
Velocity of
P relative to A
P relative to B
(3.35)
Velocity of
B relative to A
Equation (3.35) is known as the Galilean velocity transformation. It relates
the velocity of a body P with respect to frame A and its velocity with respect to
S
S
frame B (vP>A and vP>B, respectively) to the velocity of frame B with respect
S
to frame A 1vB>A2. If all three of these velocities lie along the same line, then
Eq. (3.35) reduces to Eq. (3.32) for the components of the velocities along
that line.
If the train is moving at vB>A = 3.0 m>s relative to the ground and the passenger is moving at vP>B = 1.0 m>s relative to the train, then the passenger’s velocity
3.34 (a) A passenger walking across a railroad car. (b) Position of the passenger relative to the cyclist’s frame and
S
the train’s frame. (c) Vector diagram for the velocity of the passenger relative to the ground (the cyclist’s frame), vP>A .
(a)
(b)
(c) Relative velocities
(seen from above)
B (train)
yB
S
3.0 m>s
1.0 m>s
rP>A
S
zA
rP>B
OB
Position of passenger
in both frames
xB
xA
>s
OA
rB>A
P
S
2m
= 3.
P (passenger)
S
f = 18°
v P>A
A (cyclist)
Cyclist’s
frame
vB>A
Train’s
frame
vB>A = 3.0 m>s
yA
Velocity of train
relative to cyclist
zB
vP>B = 1.0 m>s
3.5 relative Velocity
89
S
vector vP>A relative to the ground is as shown in Fig. 3.34c. The Pythagorean
theorem then gives us
vP>A = 213.0 m>s22 + 11.0 m>s22 = 210 m2>s2 = 3.2 m>s
Figure 3.34c also shows that the direction of the passenger’s velocity vector
S
relative to the ground makes an angle f with the train’s velocity vector vB>A,
where
tan f =
vP>B
vB>A
=
1.0 m>s
3.0 m>s
and
f = 18°
As in the case of motion along a straight line, we have the general rule that if
A and B are any two points or frames of reference,
S
S
vA>B = −vB>A
(3.36)
The velocity of the passenger relative to the train is the negative of the velocity of
the train relative to the passenger, and so on.
In the early 20th century Albert Einstein showed that Eq. (3.35) has to be
modified when speeds approach the speed of light, denoted by c. It turns out that
if the passenger in Fig. 3.32a could walk down the aisle at 0.30c and the train
could move at 0.90c, then her speed relative to the ground would be not 1.20c
but 0.94c; nothing can travel faster than light! We’ll return to Einstein and his
special theory of relativity in Chapter 37.
Solution
ExAmplE 3.14
flying in A crosswind
An airplane’s compass indicates that it is headed due north, and
its airspeed indicator shows that it is moving through the air at
240 km>h. If there is a 100-km>h wind from west to east, what is
the velocity of the airplane relative to the earth?
3.35 The plane is pointed north, but the wind blows east, giving
S
the resultant velocity vP>E relative to the earth.
S
vA>E = 100 km>h,
east
soLution
identify and set uP: This problem involves velocities in two
dimensions (northward and eastward), so it is a relative velocity
problem using vectors. We are given the magnitude and direction
of the velocity of the plane (P) relative to the air (A). We are also
given the magnitude and direction of the wind velocity, which is
the velocity of the air A with respect to the earth (E):
vP>A = 240 km>h
S
due north
S
due east
vA>E = 100 km>h
We’ll use Eq. (3.35) to find our target variables: the magnitude
S
and direction of velocity vP>E of the plane relative to the earth.
S
S
vP>A =
240 km>h,
north
vP>E
a
N
W
E
S
execute: From Eq. (3.35) we have
S
S
S
vP>E = vP>A + vA>E
Figure 3.35 shows that the three relative velocities constitute a
right-triangle vector addition; the unknowns are the speed vP>E
and the angle a. We find
vP>E = 21240 km>h22 + 1100 km>h22 = 260 km>h
a = arctan a
100 km>h
b = 23° E of N
240 km>h
eVaLuate: You can check the results by taking measurements on
the scale drawing in Fig. 3.35. The crosswind increases the speed
of the airplane relative to the earth, but pushes the airplane off
course.
90
ChApTer 3 Motion in Two or Three Dimensions
Solution
ExAmplE 3.15
corrEcting for A crosswind
With wind and airspeed as in Example 3.14, in what direction
should the pilot head to travel due north? What will be her velocity relative to the earth?
3.36 The pilot must point the plane in the direction of the
S
vector vP>A to travel due north relative to the earth.
S
vA>E = 100 km>h,
east
soLution
identify and set uP: Like Example 3.14, this is a relative velocity problem with vectors. Figure 3.36 is a scale drawing of the
situation. Again the vectors add in accordance with Eq. (3.35) and
form a right triangle:
S
S
S
vP>E = vP>A + vA>E
S
As Fig. 3.36 shows, the pilot points the nose of the airplane at an
angle b into the wind to compensate for the crosswind. This angle,
S
which tells us the direction of the vector vP>A (the velocity of the
airplane relative to the air), is one of our target variables. The other
target variable is the speed of the airplane over the ground, which
S
is the magnitude of the vector vP>E (the velocity of the airplane
relative to the earth). The known and unknown quantities are
vP>E = magnitude unknown
S
due north
vP>A = 240 km>h
S
direction unknown
S
due east
vA>E = 100 km>h
We’ll solve for the target variables by using Fig. 3.36 and
trigonometry.
execute: From Fig. 3.36 the speed vP>E and the angle b are
vP>E = 21240 km>h22 - 1100 km>h22 = 218 km>h
b = arcsina
100 km>h
b = 25°
240 km>h
vP>A =
240 km>h,
at angle b
S
vP>E,
north
b
N
W
E
S
The pilot should point the airplane 25° west of north, and her
ground speed is then 218 km>h.
eVaLuate: There were two target variables—the magnitude of
a vector and the direction of a vector—in both this example and
Example 3.14. In Example 3.14 the magnitude and direction referred
S
S
to the same vector 1vP>E2; here they refer to different vectors 1vP>E
S
and vP>A2.
While we expect a headwind to reduce an airplane’s speed
relative to the ground, this example shows that a crosswind does,
too. That’s an unfortunate fact of aeronautical life.
test your understanding of section 3.5 Suppose the nose of an
airplane is pointed due east and the airplane has an airspeed of 150 km>h. Due to the
wind, the airplane is moving due north relative to the ground and its speed relative to
the ground is 150 km>h. What is the velocity of the air relative to the earth? (i) 150 km>h
from east to west; (ii) 150 km>h from south to north; (iii) 150 km>h from southeast
to northwest; (iv) 212 km>h from east to west; (v) 212 km>h from south to north;
(vi) 212 km>h from southeast to northwest; (vii) there is no possible wind velocity that
could cause this. ❙
ChApTer
3 suMMary
Position, velocity, and acceleration vectors: The
S
position vector r of a point P in space is the vector
from the origin to P. Its components are the coordinates x, y, and z.
S
The average velocity vector vav during the time
S
interval ∆t is the displacement ∆r (the change in
S
position vector r ) divided by ∆t. The instantaneous
S
S
velocity vector v is the time derivative of r , and its
components are the time derivatives of x, y, and z.
S
The instantaneous speed is the magnitude of v.
S
The velocity v of a particle is always tangent to the
particle’s path. (See Example 3.1.)
S
The average acceleration vector aav during the
S
time interval ∆t equals ∆v (the change in velocity
S
vector v) divided by ∆t. The instantaneous accelS
S
eration vector a is the time derivative of v, and its
components are the time derivatives of vx , vy , and vz .
(See Example 3.2.)
The component of acceleration parallel to the direction of the instantaneous velocity affects the speed,
S
S
while the component of a perpendicular to v affects
the direction of motion. (See Examples 3.3 and 3.4.)
Projectile motion: In projectile motion with no air
resistance, ax = 0 and ay = - g. The coordinates
and velocity components are simple functions of
time, and the shape of the path is always a parabola. We usually choose the origin to be at the initial
position of the projectile. (See Examples 3.5–3.10.)
Uniform and nonuniform circular motion: When a
particle moves in a circular path of radius R with
constant speed v (uniform circular motion), its
S
acceleration a is directed toward the center of the
S
circle and perpendicular to v. The magnitude arad of
the acceleration can be expressed in terms of v and R
or in terms of R and the period T (the time for one
revolution), where v = 2pR>T. (See Examples 3.11
and 3.12.)
If the speed is not constant in circular motion
(nonuniform circular motion), there is still a radial
S
component of a given by Eq. (3.27) or (3.29), but
S
there is also a component of a parallel (tangential)
to the path. This tangential component is equal to
the rate of change of speed, dv>dt.
Relative velocity: When a body P moves relative to
a body (or reference frame) B, and B moves relative to a body (or reference frame) A, we denote the
S
velocity of P relative to B by vP>B, the velocity of P
S
relative to A by vP>A, and the velocity of B relative
S
to A by vB>A . If these velocities are all along the
same line, their components along that line are related by Eq. (3.32). More generally, these velocities
are related by Eq. (3.35). (See Examples 3.13–3.15.)
SolutionS to all exampleS
r = xnd + y ne + z kn
S
S
S
vav =
r2 − r1
∆r
=
t2 - t1
∆t
S
S
y1
S
∆r
dr
v = lim
=
∆t S 0 ∆t
dt
dx
dt
S
S
aav =
dy
dt
vy =
S
vav =
(3.2)
S
vx =
y
(3.1)
S
(3.3)
vz =
dz
dt
r1
y2
S
(3.4)
S
O
dvx
ax =
dt
dvy
ay =
dt
dvz
az =
dt
x
x2
∆x
(3.8)
S
v2
S
∆v
dv
a = lim
=
∆t S 0 ∆t
dt
(3.9)
y
S
v1
(3.10)
S
∆v
S
aav =
y = 1v0 sin a02t -
(3.19)
1 2
2 gt
S
v2
S
y
v
vy
(3.20)
vy
vx = v0 cos a0
(3.21)
S
vy = v0 sin a0 - gt
(3.22)
v2
R
S
∆v
∆t
S
v1
x
O
x = 1v0 cos a02t
arad =
r2
x1
S
S
v2 − v1
∆v
=
t2 - t1
∆t
S
arad =
S
∆r
S
∆y
S
∆r
∆t
S
v
vx
vx
v
ay = -g
vx
S
v
vy
x
O
S
v
(3.27)
4p2R
(3.29)
T2
S
v
S
arad
S
arad
S
arad
S
v
S
v
S
arad
S
arad
S
v
vP>A@x = vP>B@x + vB>A@x
(3.32)
(relative velocity along a line)
S
S
S
S
(3.35)
S
v
vB>A
vP>A
S
vP>A = vP>B + vB>A
(relative velocity in space)
S
arad
S
S
S
vP>A = vP>B + vB>A
S
vP>B
P (plane)
B (moving air)
A (ground
observer)
91
92
ChAPTer 3 Motion in Two or Three Dimensions
Solution
bridging problEm
launCHing up an inClinE
You fire a ball with an initial speed v0 at an angle f above the surface of an incline, which is itself inclined at an angle u above the
horizontal (Fig. 3.37). (a) Find the distance, measured along the
incline, from the launch point to the point when the ball strikes
the incline. (b) What angle f gives the maximum range, measured
along the incline? Ignore air resistance.
soLution guide
3. In the projectile equations in Section 3.3, the launch angle a0
is measured from the horizontal. What is this angle in terms of
u and f? What are the initial x- and y-components of the ball’s
initial velocity?
4. You’ll need to write an equation that relates x and y for points
along the incline. What is this equation? (This takes just geometry and trigonometry, not physics.)
execute
identify and set uP
1. Since there’s no air resistance, this is a problem in projectile
motion. The goal is to find the point where the ball’s parabolic
trajectory intersects the incline.
2. Choose the x- and y-axes and the position of the origin.
When in doubt, use the suggestions given in Problem-Solving
Strategy 3.1 in Section 3.3.
3.37 Launching a ball from an inclined ramp.
v0
5. Write the equations for the x-coordinate and y-coordinate of
the ball as functions of time t.
6. When the ball hits the incline, x and y are related by the equation that you found in step 4. Based on this, at what time t does
the ball hit the incline?
7. Based on your answer from step 6, at what coordinates x and y
does the ball land on the incline? How far is this point from the
launch point?
8. What value of f gives the maximum distance from the launch
point to the landing point? (Use your knowledge of calculus.)
eVaLuate
f
u
9. Check your answers for the case u = 0, which corresponds to
the incline being horizontal rather than tilted. (You already
know the answers for this case. Do you know why?)
Problems
For assigned homework and other learning materials, go to MasteringPhysics®.
., .., ...: Difficulty levels. CP: Cumulative problems incorporating material from earlier chapters. CALC: Problems requiring calculus.
DATA: Problems involving real data, scientific evidence, experimental design, and/or statistical reasoning. BIO: Biosciences problems.
discussion Questions
Q3.1 A simple pendulum (a mass swinging at the end of a string)
swings back and forth in a circular arc. What is the direction of
the acceleration of the mass when it is at the ends of the swing?
At the midpoint? In each case, explain how you obtained your
answer.
S
S
Q3.2 Redraw Fig. 3.11a if a is antiparallel to v1. Does the particle
move in a straight line? What happens to its speed?
Q3.3 A projectile moves in a parabolic path without air resistance.
S
S
S
Is there any point at which a is parallel to v? Perpendicular to v?
Explain.
Q3.4 A book slides off a horizontal tabletop. As it leaves the
table’s edge, the book has a horizontal velocity of magnitude v0.
The book strikes the floor in time t. If the initial velocity of the
book is doubled to 2v0, what happens to (a) the time the book is in
the air, (b) the horizontal distance the book travels while it is in the
air, and (c) the speed of the book just before it reaches the floor?
In particular, does each of these quantities stay the same, double,
or change in another way? Explain.
Q3.5 At the instant that you fire a bullet horizontally from a
rifle, you drop a bullet from the height of the gun barrel. If there is
no air resistance, which bullet hits the level ground first? Explain.
Q3.6 A package falls out of an airplane that is flying in a straight
line at a constant altitude and speed. If you ignore air resistance,
what would be the path of the package as observed by the pilot?
As observed by a person on the ground?
Q3.7 Sketch the six graphs of the x- and y-components of position, velocity, and acceleration versus time for projectile motion
with x0 = y0 = 0 and 0 6 a0 6 90°.
Q3.8 If a jumping frog can give itself the same initial speed regardless of the direction in which it jumps (forward or straight up),
how is the maximum vertical height to which it can jump related
to its maximum horizontal range R max = v 02>g?
Q3.9 A projectile is fired upward at an angle u above the horizontal with an initial speed v0. At its maximum height, what are
its velocity vector, its speed, and its acceleration vector?
Q3.10 In uniform circular motion, what are the average velocity
and average acceleration for one revolution? Explain.
Q3.11 In uniform circular motion, how does the acceleration
change when the speed is increased by a factor of 3? When the
radius is decreased by a factor of 2?
Q3.12 In uniform circular motion, the acceleration is perpendicular to the velocity at every instant. Is this true when the motion is
not uniform—that is, when the speed is not constant?
Q3.13 Raindrops hitting the side windows of a car in motion
often leave diagonal streaks even if there is no wind. Why? Is
the explanation the same or different for diagonal streaks on
the windshield?
exercises
Q3.14 In a rainstorm with a strong wind, what determines the
best position in which to hold an umbrella?
Q3.15 You are on the west bank of a river that is flowing north
with a speed of 1.2 m>s. Your swimming speed relative to the
water is 1.5 m>s, and the river is 60 m wide. What is your path
relative to the earth that allows you to cross the river in the shortest time? Explain your reasoning.
Q3.16 A stone is thrown into the air at an angle above the
horizontal and feels negligible air resistance. Which graph in
Fig. Q3.16 best depicts the stone’s speed v as a function of time t
while it is in the air?
Figure Q3.16
v
v
v
t
(a)
v
t
(b)
v
t
(c)
t
t
(d)
(e)
exercises
Section 3.1 Position and Velocity Vectors
.
A squirrel has x- and y-coordinates 11.1 m, 3.4 m2 at time
t1 = 0 and coordinates 15.3 m, - 0.5 m2 at time t2 = 3.0 s. For
this time interval, find (a) the components of the average velocity,
and (b) the magnitude and direction of the average velocity.
3.2 . A rhinoceros is at the origin of coordinates at time t1 = 0.
For the time interval from t1 = 0 to t2 = 12.0 s, the rhino’s average velocity has x-component - 3.8 m>s and y-component 4.9 m>s.
At time t2 = 12.0 s, (a) what are the x- and y-coordinates of the
rhino? (b) How far is the rhino from the origin?
3.3 .. CALC A web page designer creates an animation in which
a dot on a computer screen has position
3.1
r = 34.0 cm + 12.5 cm>s22t 24 nd + 15.0 cm>s2t ne .
S
(a) Find the magnitude and direction of the dot’s average velocity between t = 0 and t = 2.0 s.(b) Find the magnitude and
direction of the instantaneous velocity at t = 0, t = 1.0 s, and
t = 2.0 s. (c) Sketch the dot’s trajectory from t = 0 to t = 2.0 s,
and show the velocities calculated in part (b).
3.4 . CALC The position of a squirrel running in a park is given
S
by r = 310.280 m>s2t + 10.0360 m>s22t 24 nd + 10.0190 m>s32t 3ne .
(a) What are vx1t2 and vy1t2, the x- and y-components of the velocity of the squirrel, as functions of time? (b) At t = 5.00 s, how far
is the squirrel from its initial position? (c) At t = 5.00 s, what are
the magnitude and direction of the squirrel’s velocity?
Section 3.2 The Acceleration Vector
3.5 . A jet plane is flying at a constant altitude. At time t1 = 0,
it has components of velocity vx = 90 m>s, vy = 110 m>s. At time
t2 = 30.0 s, the components are vx = - 170 m>s, vy = 40 m>s.
(a) Sketch the velocity vectors at t1 and t2. How do these two vectors differ? For this time interval calculate (b) the components of
the average acceleration, and (c) the magnitude and direction of
the average acceleration.
3.6 .. A dog running in an open field has components of velocity vx = 2.6 m>s and vy = - 1.8 m>s at t1 = 10.0 s. For the time
interval from t1 = 10.0 s to t2 = 20.0 s, the average acceleration
of the dog has magnitude 0.45 m>s2 and direction 31.0° measured
from the + x@axis toward the + y@axis. At t2 = 20.0 s, (a) what are
93
the x- and y-components of the dog’s velocity? (b) What are the
magnitude and direction of the dog’s velocity? (c) Sketch the velocity vectors at t1 and t2. How do these two vectors differ?
3.7 .. CALC The coordinates of a bird flying in the xy-plane are
given by x1t2 = at and y1t2 = 3.0 m - bt 2, where a = 2.4 m>s
and b = 1.2 m>s2. (a) Sketch the path of the bird between t = 0
and t = 2.0 s. (b) Calculate the velocity and acceleration vectors of the bird as functions of time. (c) Calculate the magnitude
and direction of the bird’s velocity and acceleration at t = 2.0 s.
(d) Sketch the velocity and acceleration vectors at t = 2.0 s. At
this instant, is the bird’s speed increasing, decreasing, or not
changing? Is the bird turning? If so, in what direction?
3.8 . CALC A remote-controlled car is moving in a vacant parking
S
lot. The velocity of the car as a function of time is given by v =
3 2 n
2
35.00 m>s - 10.0180 m>s 2t 4 d + 32.00 m>s + 10.550 m>s 2t4 ne .
(a) What are ax1t2 and ay1t2, the x- and y-components of the car’s
velocity as functions of time? (b) What are the magnitude and
direction of the car’s velocity at t = 8.00 s? (b) What are the
magnitude and direction of the car’s acceleration at t = 8.00 s?
Section 3.3 Projectile Motion
3.9 . A physics book slides off a horizontal tabletop with a speed
of 1.10 m>s. It strikes the floor in 0.480 s. Ignore air resistance.
Find (a) the height of the tabletop above the floor; (b) the horizontal distance from the edge of the table to the point where the book
strikes the floor; (c) the horizontal and vertical components of the
book’s velocity, and the magnitude and direction of its velocity,
just before the book reaches the floor. (d) Draw x-t, y-t, vx@t, and
vy@t graphs for the motion.
3.10 .. A daring 510-N swim- Figure E3.10
mer dives off a cliff with a runv0
ning horizontal leap, as shown in
Fig. E3.10. What must her minimum speed be just as she leaves
9.00 m
the top of the cliff so that she
1.75 m
will miss the ledge at the bottom,
Ledge
which is 1.75 m wide and 9.00 m
below the top of the cliff?
. Crickets Chirpy and Milada jump from the top of a vertical cliff. Chirpy drops downward and reaches the ground in
2.70 s, while Milada jumps horizontally with an initial speed of
95.0 cm>s. How far from the base of the cliff will Milada hit the
ground? Ignore air resistance.
3.12 . A rookie quarterback throws a football with an initial
upward velocity component of 12.0 m>s and a horizontal velocity component of 20.0 m>s. Ignore air resistance. (a) How much
time is required for the football to reach the highest point of the
trajectory? (b) How high is this point? (c) How much time (after it
is thrown) is required for the football to return to its original level?
How does this compare with the time calculated in part (a)?
(d) How far has the football traveled horizontally during this
time? (e) Draw x-t, y-t, vx@t, and vy@t graphs for the motion.
3.13 .. Leaping the River I. During a storm, a car traveling
on a level horizontal road comes upon a bridge that has washed
out. The driver must get to the other side, so he decides to try leaping the river with his car. The side of the road the car is on is 21.3 m
above the river, while the opposite side is only 1.8 m above the river.
The river itself is a raging torrent 48.0 m wide. (a) How fast should
the car be traveling at the time it leaves the road in order just to
clear the river and land safely on the opposite side? (b) What is the
speed of the car just before it lands on the other side?
3.11
94
Chapter 3 Motion in two or three Dimensions
. BIO The Champion Jumper of the Insect World. The
froghopper, Philaenus spumarius, holds the world record for insect
jumps. When leaping at an angle of 58.0° above the horizontal,
some of the tiny critters have reached a maximum height of 58.7 cm
above the level ground. (See Nature, Vol. 424, July 31, 2003, p. 509.)
(a) What was the takeoff speed for such a leap? (b) What horizontal
distance did the froghopper cover for this world-record leap?
3.15 .. Inside a starship at rest on the earth, a ball rolls off the
top of a horizontal table and lands a distance D from the foot of
the table. This starship now lands on the unexplored Planet X. The
commander, Captain Curious, rolls the same ball off the same
table with the same initial speed as on earth and finds that it lands
a distance 2.76D from the foot of the table. What is the acceleration due to gravity on Planet X?
3.16 . On level ground a shell is fired with an initial velocity of
40.0 m>s at 60.0° above the horizontal and feels no appreciable air
resistance. (a) Find the horizontal and vertical components of the
shell’s initial velocity. (b) How long does it take the shell to reach
its highest point? (c) Find its maximum height above the ground.
(d) How far from its firing point does the shell land? (e) At its
highest point, find the horizontal and vertical components of its
acceleration and velocity.
3.17 . A major leaguer hits a baseball so that it leaves the bat
at a speed of 30.0 m>s and at an angle of 36.9° above the horizontal. Ignore air resistance. (a) At what two times is the baseball
at a height of 10.0 m above the point at which it left the bat?
(b) Calculate the horizontal and vertical components of the baseball’s
velocity at each of the two times calculated in part (a). (c) What
are the magnitude and direction of the baseball’s velocity when it
returns to the level at which it left the bat?
3.18 . A shot putter releases the shot some distance above the
level ground with a velocity of 12.0 m>s, 51.0° above the horizontal. The shot hits the ground 2.08 s later. Ignore air resistance.
(a) What are the components of the shot’s acceleration while in
flight? (b) What are the components of the shot’s velocity at the
beginning and at the end of its trajectory? (c) How far did she
throw the shot horizontally? (d) Why does the expression for R in
Example 3.8 not give the correct answer for part (c)? (e) How high
was the shot above the ground when she released it? (f) Draw x-t,
y-t, vx@t, and vy@t graphs for the motion.
3.19 .. Win the Prize. In a carnival booth, you can win a
stuffed giraffe if you toss a quarter into a small dish. The dish is on
a shelf above the point where the quarter leaves your hand and is
a horizontal distance of 2.1 m from this point (Fig. E3.19). If you
toss the coin with a velocity of 6.4 m>s at an angle of 60° above
the horizontal, the coin will land in the dish. Ignore air resistance.
3.14
Figure E3.19
v = 6.4 m>s
?
60°
2.1 m
(a) What is the height of the shelf above the point where the quarter leaves your hand? (b) What is the vertical component of the
velocity of the quarter just before it lands in the dish?
3.20 . Firemen use a high-pressure hose to shoot a stream of
water at a burning building. The water has a speed of 25.0 m>s
as it leaves the end of the hose and then exhibits projectile motion. The firemen adjust the angle of elevation a of the hose until
the water takes 3.00 s to reach a building 45.0 m away. Ignore air
resistance; assume that the end of the hose is at ground level.
(a) Find a. (b) Find the speed and acceleration of the water at the
highest point in its trajectory. (c) How high above the ground does
the water strike the building, and how fast is it moving just before
it hits the building?
3.21 .. A man stands on the roof of a 15.0-m-tall building and
throws a rock with a speed of 30.0 m>s at an angle of 33.0° above
the horizontal. Ignore air resistance. Calculate (a) the maximum
height above the roof that the rock reaches; (b) the speed of the
rock just before it strikes the ground; and (c) the horizontal range
from the base of the building to the point where the rock strikes
the ground. (d) Draw x-t, y-t, vx@t, and vy@t graphs for the motion.
3.22 .. A 124-kg balloon carrying a 22-kg basket is descending
with a constant downward velocity of 20.0 m>s. A 1.0-kg stone is
thrown from the basket with an initial velocity of 15.0 m>s perpendicular to the path of the descending balloon, as measured
relative to a person at rest in the basket. That person sees the stone
hit the ground 5.00 s after it was thrown. Assume that the balloon
continues its downward descent with the same constant speed of
20.0 m>s. (a) How high is the balloon when the rock is thrown?
(b) How high is the balloon when the rock hits the ground? (c) At
the instant the rock hits the ground, how far is it from the basket?
(d) Just before the rock hits the ground, find its horizontal and
vertical velocity components as measured by an observer (i) at rest
in the basket and (ii) at rest on the ground.
Section 3.4 Motion in a Circle
3.23 .. The earth has a radius of 6380 km and turns around once
on its axis in 24 h. (a) What is the radial acceleration of an object
at the earth’s equator? Give your answer in m>s2 and as a fraction
of g. (b) If arad at the equator is greater than g, objects will fly off
the earth’s surface and into space. (We will see the reason for this
in Chapter 5.) What would the period of the earth’s rotation have
to be for this to occur?
3.24 .. BIO Dizziness. Our balance is maintained, at least in
part, by the endolymph fluid in the inner ear. Spinning displaces
this fluid, causing dizziness. Suppose that a skater is spinning very
fast at 3.0 revolutions per second about a vertical axis through the
center of his head. Take the inner ear to be approximately 7.0 cm
from the axis of spin. (The distance varies from person to person.)
What is the radial acceleration (in m>s2 and in g’s) of the endolymph fluid?
3.25 . BIO Pilot Blackout in a Figure E3.25
Power Dive. A jet plane comes
in for a downward dive as shown
in Fig. E3.25. The bottom part
of the path is a quarter circle with
a radius of curvature of 280 m.
According to medical tests, pilots
will lose consciousness when
they pull out of a dive at an upward acceleration greater than
5.5g. At what speed (in m>s and in mph) will the pilot black out
during this dive?
exercises
3.26 .. A model of a helicopter rotor has four blades, each 3.40 m
long from the central shaft to the blade tip. The model is rotated in
a wind tunnel at 550 rev>min. (a) What is the linear speed of the
blade tip, in m>s? (b) What is the radial acceleration of the blade
tip expressed as a multiple of g?
3.27 . A Ferris wheel with
Figure E3.27
radius 14.0 m is turning about
a horizontal axis through its
center (Fig. E3.27). The linm
ear speed of a passenger on the
4.0
1
rim is constant and equal to
6.00 m>s. What are the magnitude and direction of the
passenger’s acceleration as she
passes through (a) the lowest
point in her circular motion
and (b) the highest point in her
circular motion? (c) How much
time does it take the Ferris wheel to make one revolution?
3.28 . The radius of the earth’s orbit around the sun (assumed
to be circular) is 1.50 * 108 km, and the earth travels around this
orbit in 365 days. (a) What is the magnitude of the orbital velocity of the earth, in m>s? (b) What is the radial acceleration of the
earth toward the sun, in m>s2 ? (c) Repeat parts (a) and (b) for the
motion of the planet Mercury (orbit radius = 5.79 * 107 km,
orbital period = 88.0 days).
3.29 .. BIO Hypergravity. At its Ames Research Center,
NASA uses its large “20-G” centrifuge to test the effects of very
large accelerations (“hypergravity”) on test pilots and astronauts.
In this device, an arm 8.84 m long rotates about one end in a
horizontal plane, and an astronaut is strapped in at the other end.
Suppose that he is aligned along the centrifuge’s arm with his
head at the outermost end. The maximum sustained acceleration
to which humans are subjected in this device is typically 12.5g.
(a) How fast must the astronaut’s head be moving to experience
this maximum acceleration? (b) What is the difference between
the acceleration of his head and feet if the astronaut is 2.00 m tall?
(c) How fast in rpm 1rev>min2 is the arm turning to produce the
maximum sustained acceleration?
Section 3.5 Relative Velocity
. A railroad flatcar is traveling to the right at a speed of
13.0 m>s relative to an observer standing on the ground. Someone
is riding a motor scooter on the flatcar (Fig. E3.30). What is the
velocity (magnitude and direction) of the scooter relative to the flatcar if the scooter’s velocity relative to the observer on the ground
is (a) 18.0 m>s to the right? (b) 3.0 m>s to the left? (c) zero?
3.30
Figure E3.30
v = 13.0 m>s
. A “moving sidewalk” in an airport terminal moves
at 1.0 m>s and is 35.0 m long. If a woman steps on at one end
and walks at 1.5 m>s relative to the moving sidewalk, how much
3.31
95
time does it take her to reach the opposite end if she walks (a) in
the same direction the sidewalk is moving? (b) In the opposite
direction?
3.32 . Two piers, A and B, are located on a river; B is 1500 m
downstream from A (Fig. E3.32). Two friends must make round
trips from pier A to pier B and return. One rows a boat at a constant speed of 4.00 km>h relative to the water; the other walks on
the shore at a constant speed of 4.00 km>h. The velocity of the
river is 2.80 km>h in the direction from A to B. How much time
does it take each person to make the round trip?
Figure E3.32
A
1500 m
B
vcurrent
3.33 .. A canoe has a velocity of 0.40 m>s southeast relative to
the earth. The canoe is on a river that is flowing 0.50 m>s east
relative to the earth. Find the velocity (magnitude and direction)
of the canoe relative to the river.
3.34 .. The nose of an ultralight plane is pointed due south, and
its airspeed indicator shows 35 m>s. The plane is in a 10@m>s wind
blowing toward the southwest relative to the earth. (a) In a vectorS
addition diagram, show the relationship of vP>E (the velocity of the
plane relative to the earth) to the two given vectors. (b) Let x be east
S
and y be north, and find the components of vP>E. (c) Find the magniS
tude and direction of vP>E.
3.35 . Crossing the River I. A river flows due south with a
speed of 2.0 m>s. You steer a motorboat across the river; your
velocity relative to the water is 4.2 m>s due east. The river is
500 m wide. (a) What is your velocity (magnitude and direction)
relative to the earth? (b) How much time is required to cross the
river? (c) How far south of your starting point will you reach the
opposite bank?
3.36 . Crossing the River II. (a) In which direction should
the motorboat in Exercise 3.35 head to reach a point on the opposite bank directly east from your starting point? (The boat’s speed
relative to the water remains 4.2 m>s.) (b) What is the velocity of
the boat relative to the earth? (c) How much time is required to
cross the river?
3.37 .. BIO Bird Migration. Canada geese migrate essentially along a north–south direction for well over a thousand kilometers in some cases, traveling at speeds up to about 100 km>h. If
one goose is flying at 100 km>h relative to the air but a 40@km>h
wind is blowing from west to east, (a) at what angle relative to
the north–south direction should this bird head to travel directly
southward relative to the ground? (b) How long will it take the
goose to cover a ground distance of 500 km from north to south?
(Note: Even on cloudy nights, many birds can navigate by using
the earth’s magnetic field to fix the north–south direction.)
3.38 .. An airplane pilot wishes to fly due west. A wind of
80.0 km>h (about 50 mi>h) is blowing toward the south. (a) If
the airspeed of the plane (its speed in still air) is 320.0 km>h
(about 200 mi>h), in which direction should the pilot head?
(b) What is the speed of the plane over the ground? Draw a vector diagram.
96
Chapter 3 Motion in two or three Dimensions
ProbLeMs
. CALC A rocket is fired at an angle from the top of a tower
of height h0 = 50.0 m. Because of the design of the engines, its
position coordinates are of the form x1t2 = A + Bt 2 and y1t2 =
C + Dt 3, where A, B, C, and D are constants. The acceleration of
S
the rocket 1.00 s after firing is a = 14.00dn + 3.00en2 m>s2. Take
the origin of coordinates to be at the base of the tower. (a) Find the
constants A, B, C, and D, including their SI units. (b) At the instant
after the rocket is fired, what are its acceleration vector and its
velocity? (c) What are the x- and y-components of the rocket’s velocity 10.0 s after it is fired, and how fast is it moving? (d) What is
the position vector of the rocket 10.0 s after it is fired?
3.40 ... CALC A faulty model rocket moves in the xy-plane
(the positive y-direction is vertically upward). The rocket’s acceleration has components ax1t2 = at 2 and ay1t2 = b - gt, where
a = 2.50 m>s4, b = 9.00 m>s2, and g = 1.40 m>s3. At t = 0
S
the rocket is at the origin and has velocity v0 = v0xnd + v0yne with
v0x = 1.00 m>s and v0y = 7.00 m>s. (a) Calculate the velocity and
position vectors as functions of time. (b) What is the maximum
height reached by the rocket? (c) What is the horizontal displacement of the rocket when it returns to y = 0?
S
3.41 .. CALC If r = bt 2nd + ct 3ne , where b and c are positive
constants, when does the velocity vector make an angle of 45.0°
with the x- and y-axes?
3.42 .. CALC The position of a dragonfly that is flying parS
allel to the ground is given as a function of time by r =
2 2 n
3 3n
32.90 m + 10.0900 m>s 2t 4 d − 10.0150 m>s 2t e . (a) At what
value of t does the velocity vector of the dragonfly make an angle
of 30.0o clockwise from the + x-axis? (b) At the time calculated in
part (a), what are the magnitude and direction of the dragonfly’s
acceleration vector?
3.43 ... CP A test rocket Figure P3.43
starting from rest at point A
is launched by accelerating
it along a 200.0-m incline at
.0 m
200
1.90 m>s2 (Fig. P3.43). The
incline rises at 35.0° above
35.0°
the horizontal, and at the inA
stant the rocket leaves it, the
engines turn off and the rocket is subject to gravity only (ignore
air resistance). Find (a) the maximum height above the ground that
the rocket reaches, and (b) the rocket’s greatest horizontal range
beyond point A.
3.44 .. CALC A bird flies in the xy-plane with a velocity vector
S
given by v = 1a - bt 22dn + gt ne , with a = 2.4 m>s, b = 1.6 m>s3,
and g = 4.0 m>s2. The positive y-direction is vertically upward.
At t = 0 the bird is at the origin. (a) Calculate the position and
acceleration vectors of the bird as functions of time. (b) What is
the bird’s altitude (y-coordinate) as it flies over x = 0 for the first
time after t = 0?
3.45 .. A sly 1.5-kg monkey and a jungle veterinarian with
a blow-gun loaded with a tranquilizer dart are 25 m above the
ground in trees 70 m apart. Just as the veterinarian shoots horizontally at the monkey, the monkey drops from the tree in a vain
attempt to escape being hit. What must the minimum muzzle velocity of the dart be for the dart to hit the monkey before the monkey reaches the ground?
3.46 ... BIO Spiraling Up. Birds of prey typically rise upward on thermals. The paths these birds take may be spiral-like.
You can model the spiral motion as uniform circular motion combined with a constant upward velocity. Assume that a bird completes a circle of radius 6.00 m every 5.00 s and rises vertically at
3.39
a constant rate of 3.00 m>s. Determine (a) the bird’s speed relative
to the ground; (b) the bird’s acceleration (magnitude and direction); and (c) the angle between the bird’s velocity vector and the
horizontal.
3.47 .. In fighting forest fires, airplanes work in support of ground
crews by dropping water on the fires. For practice, a pilot drops a
canister of red dye, hoping to hit a target on the ground below. If the
plane is flying in a horizontal path 90.0 m above the ground and has
a speed of 64.0 m>s 1143 mi>h2, at what horizontal distance from
the target should the pilot release the canister? Ignore air resistance.
3.48 ... A movie stuntwoman drops from a helicopter that is
30.0 m above the ground and moving with a constant velocity
whose components are 10.0 m>s upward and 15.0 m>s horizontal and toward the south. Ignore air resistance. (a) Where on the
ground (relative to the position of the helicopter when she drops)
should the stuntwoman have placed foam mats to break her fall?
(b) Draw x-t, y-t, vx@t, and vy@t graphs of her motion.
3.49 .. An airplane is flying with a velocity of 90.0 m>s at an
angle of 23.0° above the horizontal. When the plane is 114 m
directly above a dog that is standing on level ground, a suitcase
drops out of the luggage compartment. How far from the dog will
the suitcase land? Ignore air resistance.
3.50 .. A cannon, located 60.0 m from the base of a vertical
25.0-m-tall cliff, shoots a 15-kg shell at 43.0° above the horizontal
toward the cliff. (a) What must the minimum muzzle velocity be
for the shell to clear the top of the cliff? (b) The ground at the top
of the cliff is level, with a constant elevation of 25.0 m above the
cannon. Under the conditions of part (a), how far does the shell
land past the edge of the cliff?
3.51 . CP CALC A toy rocket is launched with an initial velocity of 12.0 m>s in the horizontal direction from the roof of a
30.0-m-tall building. The rocket’s engine produces a horizontal
acceleration of 11.60 m>s32t, in the same direction as the initial
velocity, but in the vertical direction the acceleration is g, downward. Ignore air resistance. What horizontal distance does the
rocket travel before reaching the ground?
3.52 ... An important piece of landing equipment must be
thrown to a ship, which is moving at 45.0 cm>s, before the ship
can dock. This equipment is thrown at 15.0 m>s at 60.0° above the
horizontal from the top of a tower at the edge of the water, 8.75 m
above the ship’s deck (Fig. P3.52). For this equipment to land at
the front of the ship, at what distance D from the dock should the
ship be when the equipment is thrown? Ignore air resistance.
Figure P3.52
15.0 m>s
60.0°
45.0 cm>s
8.75 m
D
3.53
...
The Longest Home Run. According to Guinness World
Records, the longest home run ever measured was hit by Roy “Dizzy”
Carlyle in a minor league game. The ball traveled 188 m (618 ft)
before landing on the ground outside the ballpark. (a) If the ball’s
Problems
initial velocity was in a direction 45° above the horizontal, what
did the initial speed of the ball need to be to produce such a home
run if the ball was hit at a point 0.9 m (3.0 ft) above ground level?
Ignore air resistance, and assume that the ground was perfectly
flat. (b) How far would the ball be above a fence 3.0 m (10 ft) high
if the fence was 116 m (380 ft) from home plate?
3.54 .. An Errand of Mercy. An airplane is dropping bales
of hay to cattle stranded in a blizzard on the Great Plains. The
pilot releases the bales at 150 m above the level ground when the
plane is flying at 75 m>s in a direction 55° above the horizontal.
How far in front of the cattle should the pilot release the hay so
that the bales land at the point where the cattle are stranded?
3.55 .. A baseball thrown at an angle of 60.0° above the horizontal strikes a building 18.0 m away at a point 8.00 m above the
point from which it is thrown. Ignore air resistance. (a) Find the
magnitude of the ball’s initial velocity (the velocity with which
the ball is thrown). (b) Find the magnitude and direction of the
velocity of the ball just before it strikes the building.
3.56 ... A water hose is used to fill a large cylindrical storage
tank of diameter D and height 2D. The hose shoots the water at
45° above the horizontal from the same level as the base of the
tank and is a distance 6D away (Fig. P3.56). For what range of
launch speeds 1v02 will the water enter the tank? Ignore air resistance, and express your answer in terms of D and g.
Figure P3.56
2D
v0 = ?
Water
45°
D
6D
3.57
..
A grasshopper leaps into the air from the edge of a vertical cliff, as shown in Fig. P3.57. Find (a) the initial speed of the
grasshopper and (b) the height of the cliff.
Figure P3.57
6.74 cm
50.0°
Not to
scale
1.06 m
3.58 .. Kicking an Extra Point. In Canadian football, after
a touchdown the team has the opportunity to earn one more point
by kicking the ball over the bar between the goal posts. The bar is
10.0 ft above the ground, and the ball is kicked from ground level,
36.0 ft horizontally from the bar (Fig. P3.58). Football regulations are stated in English units, but convert them to SI units for
this problem. (a) There is a minimum angle above the ground such
that if the ball is launched below this angle, it can never clear the
bar, no matter how fast it is kicked. What is this angle? (b) If the
ball is kicked at 45.0° above the horizontal, what must its initial
speed be if it is just to clear the bar? Express your answer in m>s
and in km>h.
97
Figure P3.58
10.0 ft
36.0 ft
3.59
...
Look Out! A snow- Figure P3.59
ball rolls off a barn roof that
slopes downward at an angle
v0 = 7.00 m>s
of 40° (Fig. P3.59). The edge
40°
of the roof is 14.0 m above the
ground, and the snowball has a
speed of 7.00 m>s as it rolls off
the roof. Ignore air resistance.
14.0 m
(a) How far from the edge of the
barn does the snowball strike
the ground if it doesn’t strike
anything else while falling?
(b) Draw x-t, y-t, vx@t, and vy@t
4.0 m
graphs for the motion in part (a).
(c) A man 1.9 m tall is standing
4.0 m from the edge of the barn.
Will the snowball hit him?
3.60 .. A boy 12.0 m above the ground in a tree throws a ball for
his dog, who is standing right below the tree and starts running
the instant the ball is thrown. If the boy throws the ball horizontally at 8.50 m>s, (a) how fast must the dog run to catch the ball
just as it reaches the ground, and (b) how far from the tree will the
dog catch the ball?
3.61 .. Suppose that the boy in Problem 3.60 throws the ball upward at 60.0° above the horizontal, but all else is the same. Repeat
parts (a) and (b) of that problem.
3.62 .. A rock is thrown with a velocity v0, at an angle of a0
from the horizontal, from the roof of a building of height h. Ignore
air resistance. Calculate the speed of the rock just before it strikes
the ground, and show that this speed is independent of a0.
3.63 .. Leaping the River II. A physics professor did daredevil stunts in his spare time. His last stunt was an attempt to jump
across a river on a motorcycle (Fig. P3.63). The takeoff ramp was
inclined at 53.0°, the river was 40.0 m wide, and the far bank was
15.0 m lower than the top of the ramp. The river itself was 100 m
below the ramp. Ignore air resistance. (a) What should his speed
have been at the top of the ramp to have just made it to the edge
of the far bank? (b) If his speed was only half the value found in
part (a), where did he land?
Figure P3.63
15.0 m
53.0°
100 m
40.0 m
98
ChAPTer 3 Motion in Two or Three Dimensions
. A 2.7-kg ball is thrown upward with an initial speed of
20.0 m>s from the edge of a 45.0-m-high cliff. At the instant the
ball is thrown, a woman starts running away from the base of
the cliff with a constant speed of 6.00 m>s. The woman runs in a
straight line on level ground. Ignore air resistance on the ball. (a)
At what angle above the horizontal should the ball be thrown so
that the runner will catch it just before it hits the ground, and how
far does she run before she catches the ball? (b) Carefully sketch
the ball’s trajectory as viewed by (i) a person at rest on the ground
and (ii) the runner.
3.65 . A 76.0-kg rock is rolling horizontally at the top of a vertical cliff that is 20 m above the surface of a lake (Fig. P3.65). The
top of the vertical face of a dam is located 100 m from the foot
of the cliff, with the top of the dam level with the surface of the
water in the lake. A level plain is 25 m below the top of the dam.
(a) What must be the minimum speed of the rock just as it leaves
the cliff so that it will reach the plain without striking the dam?
(b) How far from the foot of the dam does the rock hit the plain?
3.64
Figure P3.65
v0
20 m Cliff
100 m
Lake
Dam
25 m
Plain
3.66 .. Tossing Your Lunch. Henrietta is jogging on the sidewalk at 3.05 m>s on the way to her physics class. Bruce realizes
that she forgot her bag of bagels, so he runs to the window, which
is 38.0 m above the street level and directly above the sidewalk,
to throw the bag to her. He throws it horizontally 9.00 s after she
has passed below the window, and she catches it on the run. Ignore
air resistance. (a) With what initial speed must Bruce throw the
bagels so that Henrietta can catch the bag just before it hits the
ground? (b) Where is Henrietta when she catches the bagels?
3.67 .. A cart carrying a vertical missile launcher moves horizontally at a constant velocity of 30.0 m>s to the right. It launches
a rocket vertically upward. The missile has an initial vertical
velocity of 40.0 m>s relative to the cart. (a) How high does the
rocket go? (b) How far does the cart travel while the rocket is in
the air? (c) Where does the rocket land relative to the cart?
3.68 .. A firefighting crew uses a water cannon that shoots water
at 25.0 m>s at a fixed angle of 53.0° above the horizontal. The firefighters want to direct the water at a blaze that is 10.0 m above
ground level. How far from the building should they position
their cannon? There are two possibilities; can you get them both?
(Hint: Start with a sketch showing the trajectory of the water.)
3.69 ... In the middle of the night you are standing a horizontal
distance of 14.0 m from the high fence that surrounds the estate of
your rich uncle. The top of the fence is 5.00 m above the ground.
You have taped an important message to a rock that you want to
throw over the fence. The ground is level, and the width of the
fence is small enough to be ignored. You throw the rock from a
height of 1.60 m above the ground and at an angle of 56.0o above
the horizontal. (a) What minimum initial speed must the rock have
as it leaves your hand to clear the top of the fence? (b) For the
initial velocity calculated in part (a), what horizontal distance beyond the fence will the rock land on the ground?
3.70 ... CP Bang! A student sits atop a platform a distance h
above the ground. He throws a large firecracker horizontally with
a speed v. However, a wind blowing parallel to the ground gives
the firecracker a constant horizontal acceleration with magnitude a. As a result, the firecracker reaches the ground directly
below the student. Determine the height h in terms of v, a, and g.
Ignore the effect of air resistance on the vertical motion.
3.71 .. An airplane pilot sets a compass course due west and
maintains an airspeed of 220 km>h. After flying for 0.500 h, she
finds herself over a town 120 km west and 20 km south of her
starting point. (a) Find the wind velocity (magnitude and direction).
(b) If the wind velocity is 40 km>h due south, in what direction
should the pilot set her course to travel due west? Use the same
airspeed of 220 km>h.
3.72 .. Raindrops. When a train’s velocity is 12.0 m>s eastward, raindrops that are falling vertically with respect to the earth
make traces that are inclined 30.0° to the vertical on the windows of
the train. (a) What is the horizontal component of a drop’s velocity
with respect to the earth? With respect to the train? (b) What is
the magnitude of the velocity of the raindrop with respect to the
earth? With respect to the train?
3.73 ... In a World Cup soccer match, Juan is running due north
toward the goal with a speed of 8.00 m>s relative to the ground. A
teammate passes the ball to him. The ball has a speed of 12.0 m>s
and is moving in a direction 37.0° east of north, relative to the
ground. What are the magnitude and direction of the ball’s velocity relative to Juan?
3.74 .. An elevator is moving upward at a constant speed of
2.50 m>s. A bolt in the elevator ceiling 3.00 m above the elevator
floor works loose and falls. (a) How long does it take for the bolt to
fall to the elevator floor? What is the speed of the bolt just as it hits
the elevator floor (b) according to an observer in the elevator?
(c) According to an observer standing on one of the floor landings of
the building? (d) According to the observer in part (c), what distance
did the bolt travel between the ceiling and the floor of the elevator?
3.75 .. Two soccer players, Mia and Alice, are running as Alice
passes the ball to Mia. Mia is running due north with a speed of
6.00 m>s. The velocity of the ball relative to Mia is 5.00 m>s in a
direction 30.0o east of south. What are the magnitude and direction of the velocity of the ball relative to the ground?
3.76 .. DATA A spring-gun projects a small rock from the
ground with speed v0 at an angle u0 above the ground. You have
been asked to determine v0. From the way the spring-gun is constructed, you know that to a good approximation v0 is independent
of the launch angle. You go to a level, open field, select a launch
angle, and measure the horizontal distance the rock travels. You
use g = 9.80 m>s2 and ignore the small height of the end of the
spring-gun’s barrel above the ground. Since your measurement includes some uncertainty in values measured for the launch angle
and for the horizontal range, you repeat the measurement for several launch angles and obtain the results given in Fig. 3.76. You
Figure P3.76
Distance (m)
12.00
10.00
8.00
6.00
15.0
35.0
55.0
Launch
75.0 angle (°)
Passage Problems
ignore air resistance because there is no wind and the rock is small
and heavy. (a) Select a way to represent the data well as a straight
line. (b) Use the slope of the best straight-line fit to your data from
part (a) to calculate v0. (c) When the launch angle is 36.9o, what
maximum height above the ground does the rock reach?
3.77 .. DATA You have constructed a hair-spray-powered potato gun and want to find the muzzle speed v0 of the potatoes,
the speed they have as they leave the end of the gun barrel. You
use the same amount of hair spray each time you fire the gun, and
you have confirmed by repeated firings at the same height that
the muzzle speed is approximately the same for each firing. You
climb on a microwave relay tower (with permission, of course)
to launch the potatoes horizontally from different heights above
the ground. Your friend measures the height of the gun barrel
above the ground and the range R of each potato. You obtain the
following data:
Launch height h
Horizontal range R
2.00 m
6.00 m
9.00 m
12.00 m
10.4 m
17.1 m
21.3 m
25.8 m
Each of the values of h and R has some measurement error: The
muzzle speed is not precisely the same each time, and the barrel
isn’t precisely horizontal. So you use all of the measurements to
get the best estimate of v0. No wind is blowing, so you decide
to ignore air resistance. You use g = 9.80 m>s2 in your analysis. (a) Select a way to represent the data well as a straight line.
(b) Use the slope of the best-fit line from part (a) to calculate the
average value of v0. (c) What would be the horizontal range of a
potato that is fired from ground level at an angle of 30.0o above
the horizontal? Use the value of v0 that you calculated in part (b).
3.78 ... DATA You are a member of a geological team in
Central Africa. Your team comes upon a wide river that is flowing
east. You must determine the width of the river and the current
speed (the speed of the water relative to the earth). You have a
small boat with an outboard motor. By measuring the time it takes
to cross a pond where the water isn’t flowing, you have calibrated
the throttle settings to the speed of the boat in still water. You
set the throttle so that the speed of the boat relative to the river is a
constant 6.00 m>s. Traveling due north across the river, you reach
the opposite bank in 20.1 s. For the return trip, you change the
throttle setting so that the speed of the boat relative to the water
is 9.00 m>s. You travel due south from one bank to the other and
cross the river in 11.2 s. (a) How wide is the river, and what is the
current speed? (b) With the throttle set so that the speed of the
boat relative to the water is 6.00 m>s, what is the shortest time in
which you could cross the river, and where on the far bank would
you land?
chaLLenge ProbLeMs
... CALC A projectile thrown from a point P moves in such
a way that its distance from P is always increasing. Find the maximum angle above the horizontal with which the projectile could
have been thrown. Ignore air resistance.
3.80 ... Two students are canoeing on a river. While heading
upstream, they accidentally drop an empty bottle overboard. They
then continue paddling for 60 minutes, reaching a point 2.0 km
farther upstream. At this point they realize that the bottle is
missing and, driven by ecological awareness, they turn around
3.79
99
and head downstream. They catch up with and retrieve the bottle
(which has been moving along with the current) 5.0 km downstream
from the turnaround point. (a) Assuming a constant paddling effort
throughout, how fast is the river flowing? (b) What would the canoe
speed in a still lake be for the same paddling effort?
3.81 ... CP A rocket designed to place small payloads into orbit
is carried to an altitude of 12.0 km above sea level by a converted
airliner. When the airliner is flying in a straight line at a constant
speed of 850 km>h, the rocket is dropped. After the drop, the airliner maintains the same altitude and speed and continues to fly
in a straight line. The rocket falls for a brief time, after which its
rocket motor turns on. Once that motor is on, the combined effects
of thrust and gravity give the rocket a constant acceleration of
magnitude 3.00g directed at an angle of 30.0° above the horizontal. For safety, the rocket should be at least 1.00 km in front of the
airliner when it climbs through the airliner’s altitude. Your job is
to determine the minimum time that the rocket must fall before
its engine starts. Ignore air resistance. Your answer should include
(i) a diagram showing the flight paths of both the rocket and
the airliner, labeled at several points with vectors for their velocities and accelerations; (ii) an x-t graph showing the motions of
both the rocket and the airliner; and (iii) a y-t graph showing the
motions of both the rocket and the airliner. In the diagram and
the graphs, indicate when the rocket is dropped, when the rocket
motor turns on, and when the rocket climbs through the altitude
of the airliner.
Passage ProbLeMs
BIO BallisTic sEEd disPErsal. Some plants disperse
their seeds when the fruit splits and contracts, propelling the
seeds through the air. The trajectory of these seeds can be determined with a high-speed camera. In an experiment on one type
of plant, seeds are projected at 20 cm above ground level with
initial speeds between 2.3 m>s and 4.6 m>s. The launch angle is
measured from the horizontal, with + 90° corresponding to an
initial velocity straight up and -90° straight down.
3.82 The experiment is designed so that the seeds move no
more than 0.20 mm between photographic frames. What minimum frame rate for the high-speed camera is needed to achieve
this? (a) 250 frames>s; (b) 2500 frames>s; (c) 25,000 frames>s;
(d) 250,000 frames>s.
3.83 About how long does it take a seed launched at 90° at the
highest possible initial speed to reach its maximum height? Ignore
air resistance. (a) 0.23 s; (b) 0.47 s; (c) 1.0 s; (d) 2.3 s.
3.84 If a seed is launched at an angle of 0° with the maximum
initial speed, how far from the plant will it land? Ignore air resistance, and assume that the ground is flat. (a) 20 cm; (b) 93 cm;
(c) 2.2 m; (d) 4.6 m.
3.85 A large number of seeds are observed, and their initial
launch angles are recorded. The range of projection angles is
found to be -51° to 75°, with a mean of 31°. Approximately 65%
of the seeds are launched between 6° and 56°. (See W. J. Garrison
et al., “Ballistic seed projection in two herbaceous species,” Amer.
J. Bot., Sept. 2000, 87:9, 1257–64.) Which of these hypotheses is
best supported by the data? Seeds are preferentially launched
(a) at angles that maximize the height they travel above the plant;
(b) at angles below the horizontal in order to drive the seeds into
the ground with more force; (c) at angles that maximize the horizontal distance the seeds travel from the plant; (d) at angles that
minimize the time the seeds spend exposed to the air.
100
Chapter 3 Motion in two or three Dimensions
Answers
chapter opening Question
?
(iii) A cyclist going around a curve at constant speed has an acceleration directed toward the inside of the curve (see Section 3.2,
especially Fig. 3.12a).
test your understanding Questions
S
3.1 (iii) If the instantaneous velocity v is constant over an interval, its value at any point (including the end of the interval)
S
is the same as the average velocity vav over the interval. In (i)
S
and (ii) the direction of v at the end of the interval is tangent to
S
the path at that point, while the direction of vav points from the
beginning of the path to its end (in the direction of the net disS
S
placement). In (iv) both v and vav are directed along the straight
S
line, but v has a greater magnitude because the speed has been
increasing.
3.2 Vector 7 At the high point of the sled’s path, the speed is
minimum. At that point the speed is neither increasing nor decreasing, and the parallel component of the acceleration (that is,
the horizontal component) is zero. The acceleration has only a
perpendicular component toward the inside of the sled’s curved
path. In other words, the acceleration is downward.
3.3 (i) If there were no gravity 1g = 02, the monkey would not
fall and the dart would follow a straight-line path (shown as a
dashed line). The effect of gravity is to make both the monkey
and the dart fall the same distance 12 gt 2 below their g = 0 positions. Point A is the same distance below the monkey’s initial
position as point P is below the dashed straight line, so point A is
where we would find the monkey at the time in question.
3.4 (ii) At both the top and bottom of the loop, the acceleration
is purely radial and is given by Eq. (3.27). Radius R is the same
at both points, so the difference in acceleration is due purely to
differences in speed. Since arad is proportional to the square of v,
the speed must be twice as great at the bottom of the loop as at
the top.
3.5 (vi) The effect of the wind is to cancel the airplane’s eastward motion and give it a northward motion. So the velocity of
the air relative to the ground (the wind velocity) must have one
150-km>h component to the west and one 150-km>h component
to the north. The combination of these is a vector of magnitude
21150 km>h22 + 1150 km>h22 = 212 km>h that points to the
northwest.
bridging Problem
(a) R =
2v 02 cos1u + f2sin f
g
cos2 u
(b) f = 45° -
u
2
?
Under what circumstances
does the barbell push on
the weightlifter just as hard as
he pushes on the barbell?
(i) When he holds the barbell
stationary; (ii) when he raises
the barbell; (iii) when he lowers
the barbell; (iv) two of (i), (ii),
and (iii); (v) all of (i), (ii), and (iii);
(vi) none of these.
4
NewtoN’s Laws
of MotioN
Learning goaLs
Looking forward at …
4.1 What the concept of force means in physics,
4.2
4.3
4.4
4.5
4.6
why forces are vectors, and the significance
of the net force on an object.
What happens when the net force on an
object is zero, and the significance of inertial
frames of reference.
How the acceleration of an object is determined by the net force on the object and
the object’s mass.
The difference between the mass of an
object and its weight.
How the forces that two objects exert on
each other are related.
How to use a free-body diagram to help
analyze the forces on an object.
Looking back at …
2.4 Straight-line motion with constant
acceleration.
2.5 The motion of freely falling bodies.
3.2 Acceleration as a vector.
3.4 Uniform circular motion.
3.5 Relative velocity.
W
e’ve seen in the last two chapters how to use kinematics to describe
motion in one, two, or three dimensions. But what causes bodies to
move the way that they do? For example, why does a dropped feather
fall more slowly than a dropped baseball? Why do you feel pushed backward in a
car that accelerates forward? The answers to such questions take us into the subject of dynamics, the relationship of motion to the forces that cause it.
The principles of dynamics were clearly stated for the first time by Sir Isaac
Newton (1642–1727); today we call them Newton’s laws of motion. The first law
states that when the net force on a body is zero, its motion doesn’t change. The
second law tells us that a body accelerates when the net force is not zero. The
third law relates the forces that two interacting bodies exert on each other.
Newton did not derive the three laws of motion, but rather deduced them from
a multitude of experiments performed by other scientists, especially Galileo Galilei
(who died the year Newton was born). Newton’s laws are the foundation of
classical mechanics (also called Newtonian mechanics); using them, we can
understand most familiar kinds of motion. Newton’s laws need modification only
for situations involving extremely high speeds (near the speed of light) or very
small sizes (such as within the atom).
Newton’s laws are very simple to state, yet many students find these laws
difficult to grasp and to work with. The reason is that before studying physics,
you’ve spent years walking, throwing balls, pushing boxes, and doing dozens of
things that involve motion. Along the way, you’ve developed a set of “common
sense” ideas about motion and its causes. But many of these “common sense”
ideas don’t stand up to logical analysis. A big part of the job of this chapter—
and of the rest of our study of physics—is helping you recognize how “common
sense” ideas can sometimes lead you astray, and how to adjust your understanding of the physical world to make it consistent with what experiments tell us.
101
102
Chapter 4 Newton’s Laws of Motion
4.1 Force and inTeracTions
4.1 Some properties of forces.
• A force is a push or a pull.
• A force is an interaction between two objects
or between an object and its environment.
• A force is a vector quantity, with magnitude
and direction.
S
F (force)
S
F
Push
Pull
4.2 Four common types of forces.
S
(a) Normal force n: When an object rests or
pushes on a surface, the surface exerts a push on
it that is directed perpendicular to the surface.
S
n
S
n
S
(b) Friction force f: In addition to the normal
force, a surface may exert a friction force on an
object, directed parallel to the surface.
S
n
S
f
S
(c) Tension force T: A pulling force exerted on
an object by a rope, cord, etc.
S
T
S
(d) Weight w: The pull of gravity on an object
is a long-range force (a force that acts over
a distance).
S
w
In everyday language, a force is a push or a pull. A better definition is that a
force is an interaction between two bodies or between a body and its environment (Fig. 4.1). That’s why we always refer to the force that one body exerts on a
second body. When you push on a car that is stuck in the snow, you exert a force
on the car; a steel cable exerts a force on the beam it is hoisting at a construction
site; and so on. As Fig. 4.1 shows, force is a vector quantity; you can push or pull
a body in different directions.
When a force involves direct contact between two bodies, such as a push
or pull that you exert on an object with your hand, we call it a contact force.
Figures 4.2a, 4.2b, and 4.2c show three common types of contact forces. The
normal force (Fig. 4.2a) is exerted on an object by any surface with which it is
in contact. The adjective normal means that the force always acts perpendicular
to the surface of contact, no matter what the angle of that surface. By contrast,
the friction force (Fig. 4.2b) exerted on an object by a surface acts parallel to
the surface, in the direction that opposes sliding. The pulling force exerted by a
stretched rope or cord on an object to which it’s attached is called a tension force
(Fig. 4.2c). When you tug on your dog’s leash, the force that pulls on her collar is
a tension force.
In addition to contact forces, there are long-range forces that act even when
the bodies are separated by empty space. The force between two magnets is an
example of a long-range force, as is the force of gravity (Fig. 4.2d); the earth
pulls a dropped object toward it even though there is no direct contact between
the object and the earth. The gravitational force that the earth exerts on your
body is called your weight. S
To describe a force vector F, we need to describe the direction in which it acts
as well as its magnitude, the quantity that describes “how much” or “how hard”
the force pushes or pulls. The SI unit of the magnitude of force is the newton,
abbreviated N. (We’ll give a precise definition of the newton in Section 4.3.)
Table 4.1 lists some typical force magnitudes.
A common instrument for measuring force magnitudes is the spring balance.
It consists of a coil spring enclosed in a case with a pointer attached to one end.
When forces are applied to the ends of the spring, it stretches by an amount that
depends on the force. We can make a scale for the pointer by using a number of
identical bodies with weights of exactly 1 N each. When one, two, or more of
these are suspended simultaneously from the balance, the total force stretching
the spring is 1 N, 2 N, and so on, and we can label the corresponding positions
of the pointer 1 N, 2 N, and so on. Then we can use this instrument to measure
the magnitude of an unknown force. We can also make a similar instrument that
measures pushes instead of pulls.
Table 4.1 Typical Force Magnitudes
Sun’s gravitational force on the earth
3.5 * 1022 N
Weight of a large blue whale
1.9 * 106 N
Maximum pulling force of a locomotive
8.9 * 105 N
Weight of a 250-lb linebacker
1.1 * 103 N
Weight of a medium apple
1N
Weight of the smallest insect eggs
2 * 10-6 N
Electric attraction between the proton and the electron in a hydrogen atom
8.2 * 10-8 N
Weight of a very small bacterium
1 * 10-18 N
Weight of a hydrogen atom
1.6 * 10-26 N
Weight of an electron
8.9 * 10-30 N
Gravitational attraction between the proton and the electron in a hydrogen atom
3.6 * 10-47 N
4.1 Force and Interactions
103
4.3 Using a vector arrow to denote the force that we exert when (a) pulling a block with
a string or (b) pushing a block with a stick.
(a) A 10-N pull directed 30° above
the horizontal
10 N
(b) A 10-N push directed 45° below
the horizontal
10 N
30°
45°
Figure 4.3 shows a spring balance being used to measure a pull or push that
we apply to a box. In each case we draw a vector to represent the applied force.
The length of the vector shows the magnitude; the longer the vector, the greater
the force magnitude.
superposition of Forces
When you throw a ball, at least two forces act on it: the push of your Shand and
the
S
downward pull of gravity. Experiment shows that when two forces F1 and F2 act
at the same time at the same point on a body
(Fig. 4.4), the effect on the body’s
S
motion is the same as if a single force
R
were
acting equal to the vector sum,
S
S
S
or resultant, of the original forces: R = F1 + F2. More generally, any number of
forces applied at a point on a body have the same effect as a single force equal
to the vector sum of the forces. This important principle is called superposition
of forces.
Since forces are vector quantities and add like vectors, we can use all of the
rules of vector mathematics that we learned in Chapter 1 to solve problems that
involve vectors. This would be a good time to review the rules for vector addition
presented in Sections 1.7 and 1.8.
We learned in Section 1.8 that it’s Seasiest to add vectors by using components.
That’s why we often describe a force F in terms of its x- and y-components Fx and
Fy. Note that the x- and y-coordinate axes do not have to be horizontal and vertical, respectively.
As an example, Fig. 4.5 shows a crate being pulled up a ramp by
S
a force F. In this situation it’s most convenient to choose one axis to be parallel
to the ramp and the other to be perpendicular to the ramp. For the case shown in
Fig. 4.5, both Fx and Fy are positive; in other
situations, depending on your choice
S
of axes and the orientation of the force F, either Fx or Fy may be negative or zero.
line in force diagrams In Fig. 4.5 we draw a wiggly line
cauTion Using a wiggly
S
through the force vector F to show that we have replaced it by its x- and y-components.
Otherwise, the diagram would include the same force twice. We will draw such a wiggly
line in any force diagram where a force is replaced by its components. Look for this wiggly line in other figures in this and subsequent chapters. ❙
We will often need to find the vector sum (resultant) of all forces acting on a
body. We call this the net force acting on the body. We will use the Greek letter
g (capital sigma, equivalent
to Sthe Roman S) as a shorthand notation for a sum.
S
S
If the forces are labeled F1, F2, F3, and so on, we can write
S
S
S
S
S
The net force
c
R
=
gF
=
F
+
F
+
F
1
2
3 +
acting on a body ...
... is the vector sum, or resultant, of all individual forces acting on that body.
4.4 Superposition of forces.
S
S
Two forces F1 and F2 acting on a body at
point
O have the same effect as a single force
S
R equal to their vector sum.
S
F2
S
R
O
S
F1
S
4.5 Fx and Fy are the components of F
parallel and perpendicular to the sloping
surface of the inclined plane.
We cross out a
vector when we
replace it by its
components.
x
y
S
F
u
Fy
Fx
O
(4.1)
The x- and y-axes can have any orientation,
just so they’re mutually perpendicular.
104
Chapter 4 Newton’s Laws of Motion
We read g F as “the vector sum of the forces” or “the net force.” The
x-component of the net force is the sum of the x-components of the individual
forces, and likewise for the y-component (Fig. 4.6):
S
4.6 Finding theScomponents of Sthe vector
S
sum (resultant) R of two forces F1 and F2.
S
The y-component of R
equals the sum of
the y-S
S
components of F1 and F2.
y
S
S
S
F2
Ry
S
F1
F1x
O
F2x
(4.2)
R = 2Rx2 + Ry2
and the angle u between R and the +x@axis can be found from the relationship
tan u = Ry>Rx . The components Rx and Ry may be positive, negative, or zero, and
the angle u may be in any of the four quadrants.
In three-dimensional problems, forces may also have z-components; then we
add the equation Rz = g Fz to Eq. (4.2). The magnitude of the net force is then
S
F1y
Ry = g Fy
Each component may be positive or negative, so be careful with signs when you
evaluate these sums.
Once
we have
Rx and Ry we can find the magnitude and direction of the net
S
S
force R = g F acting on the body. The magnitude is
R = ΣF
F2y
Rx = g Fx
The same goes for
the x-components.
x
Rx
R = 2Rx2 + Ry2 + Rz2
Solution
ExamplE 4.1 SupErpoSition of forcES
Three professional wrestlers are fighting over a champion’s
belt. Figure 4.7a shows the horizontal force each wrestler
applies to the belt, as viewed from above. The forces have magnitudes F1 = 250 N, F2 = 50 N, and F3 = 120 N. Find the x- and
y-components of the net force on the belt, and find its magnitude
and direction.
Solution
identify and Set up: This is a problem in vector addition in
which the vectors happen to represent forces.
We want to find
S
the x- and y-components of the net force R, so we’ll use the component method of vector addition
expressed by Eqs. (4.2). Once
S
we know the components of R, we can find its magnitude and
direction.
S
execute: From Fig. 4.7a the angles between the three forces F1,
S
S
F2, and F3 and the + x-axis are u1 = 180° - 53° = 127°, u2 = 0°,
and u3 = 270°. The x- and y-components of the three forces are
F1x = 1250 N2 cos 127° = - 150 N
4.7 (a) Three forces acting on a belt. (b) The net force R = gF
S
S
and its components.
S
F1y
x- and
y-components
S
of F1
Net
force
S
S
R = ΣF
Ry
u = 141°
S
F2
x
S
F2 has zero
y-component.
F1x
S
F3 has zero
x-component.
y
(b)
53°
S
F3
F3x = 1120 N2 cos 270° = 0 N
F3y = 1120 N2 sin 270° = -120 N
From Eqs. (4.2) the net force R = gF has components
S
S
Rx = F1x + F2x + F3x = 1- 150 N2 + 50 N + 0 N = - 100 N
Ry = F1y + F2y + F3y = 200 N + 0 N + 1-120 N2 = 80 N
The net force has a negative x-component and a positive y-component, as Fig. 4.7b shows.
S
The magnitude of R is
R = 2Rx2 + Ry2 = 21-100 N22 + 180 N22 = 128 N
To find the angle between the net force and the + x-axis, we use
Eq. (1.7):
Ry
Rx
= arctan a
80 N
b = arctan 1- 0.802
-100 N
The arctangent of -0.80 is -39°, but Fig. 4.7b shows that the net
force lies in the second quadrant. Hence the correct solution is
u = -39° + 180° = 141°.
evaluate: The net force is not zero. Your intuition should sug-
y
F1
F2y = 150 N2 sin 0° = 0 N
u = arctan
F1y = 1250 N2 sin 127° = 200 N
(a)
F2x = 150 N2 cos 0° = 50 N
Rx
x
gest that wrestler 1 (who exerts the greatest force on the belt,
F1 = 250 N) will walk away with it whenS the struggle ends.
You
shouldS check the direction of R by adding the vectors
S
S
F1, F2, and F3 graphically. Does your drawing show that
S
S
S
S
R = F1 + F2 + F3 points in the second quadrant as we found?
4.2 Newton’s First Law
105
S
TesT your undersTanding oF secTion 4.1 Figure 4.5 shows a force F
acting on a crate. With the x- and y-axes shown in the figure, which statement about
the components of the gravitational force that the earth exerts on the crate (the crate’s
weight) is correct? (i) Both the x- and y-components are positive. (ii) The x-component
is zero and the y-component is positive. (iii) The x-component is negative and the
y-component is positive. (iv) Both the x- and y-components are negative. (v) The
x-component is zero and the y-component is negative. (vi) The x-component is positive
and the y-component is negative. ❙
4.2 newTon’s FirsT Law
How do the forces that act on a body affect its motion? To begin to answer this
question, let’s first consider what happens when the net force on a body is zero.
You would almost certainly agree that if a body is at rest, and if no net force acts
on it (that is, no net push or pull), that body will remain at rest. But what if there
is zero net force acting on a body in motion?
To see what happens in this case, suppose you slide a hockey puck along a
horizontal tabletop, applying a horizontal force to it with your hand (Fig. 4.8a).
After you stop pushing, the puck does not continue to move indefinitely; it slows
down and stops. To keep it moving, you have to keep pushing (that is, applying a
force). You might come to the “common sense” conclusion that bodies in motion
naturally come to rest and that a force is required to sustain motion.
But now imagine pushing the puck across a smooth surface of ice (Fig. 4.8b).
After you quit pushing, the puck will slide a lot farther before it stops. Put it on
an air-hockey table, where it floats on a thin cushion of air, and it moves still
farther (Fig. 4.8c). In each case, what slows the puck down is friction, an interaction between the lower surface of the puck and the surface on which it slides.
Each surface exerts a friction force on the puck that resists the puck’s motion; the
difference in the three cases is the magnitude of the friction force. The ice exerts
less friction than the tabletop, so the puck travels farther. The gas molecules of
the air-hockey table exert the least friction of all. If we could eliminate friction
completely, the puck would never slow down, and we would need no force at all
to keep the puck moving once it had been started. Thus the “common sense” idea
that a force is required to sustain motion is incorrect.
Experiments like the ones we’ve just described show that when no net force
acts on a body, the body either remains at rest or moves with constant velocity in
a straight line. Once a body has been set in motion, no net force is needed to keep
it moving. We call this observation Newton’s first law of motion:
newTon’s FirsT Law oF MoTion: A body acted on by no net force has a
constant velocity (which may be zero) and zero acceleration.
The tendency of a body to keep moving once it is set in motion is called inertia.
You use inertia when you try to get ketchup out of a bottle by shaking it. First
you start the bottle (and the ketchup inside) moving forward; when you jerk the
bottle back, the ketchup tends to keep moving forward and, you hope, ends up on
your burger. Inertia is also the tendency of a body at rest to remain at rest. You
may have seen a tablecloth yanked out from under the china without breaking
anything. The force on the china isn’t great enough to make it move appreciably
during the short time it takes to pull the tablecloth away.
It’s important to note that the net force is what matters in Newton’s first law.
For example, a physics book at rest on a horizontal tabletop has two forces acting on it: an upward supporting force, or normal force, exerted by the tabletop
(see Fig. 4.2a) and the downward force of the earth’s gravity (which acts even if
the tabletop is elevated above the ground; see Fig. 4.2d). The upward push of the
surface is just as great as the downward pull of gravity, so the net force acting
4.8 The slicker the surface, the farther a
puck slides after being given an initial
velocity. On an air-hockey table (c) the
friction force is practically zero, so the
puck continues with almost constant
velocity.
(a) Table: puck stops short.
(b) Ice: puck slides farther.
(c) Air-hockey table: puck slides even farther.
.................
................
...............
...............
..............
. . . . . . . . . . . . . .
. . . . . .
. . . . .
. . . . .
. . . . .
106
Chapter 4 Newton’s Laws of Motion
4.9 (a) A hockey puck accelerates
in the
S
direction of a net applied force F1 . (b) When
the net force is zero, the acceleration is zero,
and the puck is in equilibrium.
(a) A puck on a frictionless surface
accelerates when acted on by a
single horizontal force.
(b) This puck is acted on by two
horizontal forces whose vector sum
is zero. The puck behaves as though
no forces act on it.
S
ΣF = 0
S
a=0
S
a
S
S
F1
sledding with newton’s
First Law The downward force of gravity
application
acting on the child and sled is balanced by an
upward normal force exerted by the ground.
The adult’s foot exerts a forward force that
balances the backward force of friction on the
sled. Hence there is no net force on the child
and sled, and they slide with a constant
velocity.
S
F1
F2
on the book (that is, the vector sum of the two forces) is zero. In agreement with
Newton’s first law, if the book is at rest on the tabletop, it remains at rest. The
same principle applies to a hockey puck sliding on a horizontal, frictionless
surface: The vector sum of the upward push of the surface and the downward
pull of gravity is zero. Once the puck is in motion, it continues to move with
constant velocity because the net force acting on it is zero.
Here’s another example. Suppose a hockey puck rests on a horizontal surface
with negligible friction, such as an air-hockey table Sor a slab of wet ice. If the
puck is initially at rest and a single horizontal force F1 acts on it (Fig. 4.9a), the
puck starts to move. If the puck is in motion to begin with, the force changes its
speed, its direction, or both,
depending on the direction of the force. In this case
S
the net force is equal to F1, which is not zero. (There are also two vertical forces:
the earth’s gravitational attraction and the upward normal force exerted by the
surface. But as we mentioned earlier, theseStwo forces cancel.)
S
Now suppose we apply a second force, F2 (Fig. 4.9b), equal in magnitude
to SF1
S
but opposite in direction. The two forces are negatives of each other, F2 = −F1,
and their vector sum is zero:
gF = F1 + F2 = F1 + 1 −F1 2 = 0
S
S
S
S
S
Again, we find that if the body is at rest at the start, it remains at rest; if it is
initially moving, it continues to move in the same direction with constant speed.
These results show that in Newton’s first law, zero net force is equivalent to no
force at all. This is just the principle of superposition of forces that we saw in
Section 4.1.
When a body is either at rest or moving with constant velocity (in a straight
line with constant speed), we say that the body is in equilibrium. For a body to
be in equilibrium, it must be acted on by no forces, or by several forces such that
their vector sum—that is, the net force—is zero:
Newton’s first law:
Net force on a body ...
gF = 0
S
... must be zero if body
is in equilibrium.
(4.3)
We’re assuming that the body can be represented adequately as a point particle. When the body has finite size, we also have to consider where on the body
the forces are applied. We’ll return to this point in Chapter 11.
ZEro nEt forcE mEanS conStant vElocity
In the classic 1950 science-fiction film Rocketship X-M, a spaceship is moving in the vacuum of outer space, far from any star or
planet, when its engine dies. As a result, the spaceship slows down
and stops. What does Newton’s first law say about this scene?
Solution
concEptUal ExamplE 4.2
soLuTion
No forces act on the spaceship after the engine dies, so according
to Newton’s first law it will not stop but will continue to move in a
straight line with constant speed. Some science-fiction movies are
based on accurate science; this is not one of them.
4.2 Newton’s First Law
Solution
concEptUal ExamplE 4.3
107
conStant vElocity mEanS ZEro nEt forcE
You are driving a Maserati GranTurismo S on a straight testing track
at a constant speed of 250 km>h. You pass a 1971 Volkswagen Beetle
doing a constant 75 km>h. On which car is the net force greater?
Maserati’s high-power engine, it’s true that the track exerts
a greater forward force on your Maserati than it does on the
Volkswagen. But a backward force also acts on each car due to
road friction and air resistance. When the car is traveling with constant velocity, the vector sum of the forward and backward forces
is zero. There is more air resistance on the fast-moving Maserati
than on the slow-moving Volkswagen, which is why the Maserati’s
engine must be more powerful than that of the Volkswagen.
soLuTion
The key word in this question is “net.” Both cars are in equilibrium
because their velocities are constant; Newton’s first law therefore
says that the net force on each car is zero.
This seems to contradict the “common sense” idea that the
faster car must have a greater force pushing it. Thanks to your
inertial Frames of reference
In discussing relative velocity in Section 3.5, we introduced the concept of frame of
reference. This concept is central to Newton’s laws of motion. Suppose you are in
a bus that is traveling on a straight road and speeding up. If you could stand in the
aisle on roller skates, you would start moving backward relative to the bus as the
bus gains speed. If instead the bus was slowing to a stop, you would start moving
forward down the aisle. In either case, it looks as though Newton’s first law is not
obeyed; there is no net force acting on you, yet your velocity changes. What’s wrong?
The point is that the bus is accelerating with respect to the earth and is not a
suitable frame of reference for Newton’s first law. This law is valid in some frames
of reference and not valid in others. A frame of reference in which Newton’s first
law is valid is called an inertial frame of reference. The earth is at least approximately an inertial frame of reference, but the bus is not. (The earth is not a completely inertial frame, owing to the acceleration associated with its rotation and its
motion around the sun. These effects are quite small, however; see Exercises 3.23
and 3.28.) Because Newton’s first law is used to define what we mean by an inertial
frame of reference, it is sometimes called the law of inertia.
Figure 4.10 helps us understand what you experience when riding in a vehicle
that’s accelerating. In Fig. 4.10a, a vehicle is initially at rest and then begins to
4.10 Riding in an accelerating vehicle.
(a)
(b)
Initially, you and the
vehicle are at rest.
S
v
S
t = 0
v=0
t = 0
t = 0
S
a
S
a
S
a
S
v
S
t = ∆t
t = ∆t
v
S
a
S
a
S
v
t = 2∆t
S
a
t = ∆t
v
S
t = 2∆t
(c) The vehicle rounds a turn
at constant speed.
Initially, you and the
vehicle are in motion.
S
a
S
a
S
v
S
v
t = 3∆t
S
a
You tend to remain at rest as the
vehicle accelerates around you.
t = 3∆t
S
a
t = 2∆t
S
a
You tend to continue moving
with constant velocity as the
vehicle slows down around you.
S
v
You tend to continue moving in a
straight line as the vehicle turns.
108
Chapter 4 Newton’s Laws of Motion
4.11 From the frame of reference of the
car, it seems as though a force is pushing
the crash test dummies forward as the car
comes to a sudden stop. But there is really
no such force: As the car stops, the dummies keep moving forward as a consequence of Newton’s first law.
accelerate to the right. A passenger standing on roller skates (which nearly eliminate the effects of friction) has virtually no net force acting on her, so she tends
to remain at rest relative to the inertial frame of the earth. As the vehicle accelerates around her, she moves backward relative to the vehicle. In the same way,
a passenger in a vehicle that is slowing down tends to continue moving with
constant velocity relative to the earth, and so moves forward relative to the
vehicle (Fig. 4.10b). A vehicle is also accelerating if it moves at a constant speed
but is turning (Fig. 4.10c). In this case a passenger tends to continue moving relative to the earth at constant speed in a straight line; relative to the vehicle, the
passenger moves to the side of the vehicle on the outside of the turn.
In each case shown in Fig. 4.10, an observer in the vehicle’s frame of reference
might be tempted to conclude that there is a net force acting on the passenger,
since the passenger’s velocity relative to the vehicle changes in each case. This
conclusion is simply wrong; the net force on the passenger is indeed zero. The
vehicle observer’s mistake is in trying to apply Newton’s first law in the vehicle’s
frame of reference, which is not an inertial frame and in which Newton’s first law
isn’t valid (Fig. 4.11). In this book we will use only inertial frames of reference.
We’ve mentioned only one (approximately) inertial frame of reference: the
earth’s surface. But there are many inertial frames. If we have an inertial frame
of reference A, in which Newton’s first law is obeyed, then any second frame
of reference B will also be inertial if it moves relative to A with constant
S
velocity v B>A . We can prove this by using the relative-velocity relationship
Eq. (3.35) from Section 3.5:
S
S
S
vP>A = vP>B + vB>A
S
Suppose that P is a body that moves with constant velocity v P>A with respect
to an inertial frame A. By Newton’s first law the net force on this body is zero.
S
The velocity of P relative to another frame B has a different value, v P>B =
S
S
S
v P>A − v B>A . But if the relative velocity v B>A of the two frames is constant,
S
then v P>B is constant as well. Thus B is also an inertial frame; the velocity of P
in this frame is constant, and the net force on P is zero, so Newton’s first law
is obeyed in B. Observers in frames A and B will disagree about the velocity
of P, but they will agree that P has a constant velocity (zero acceleration) and
has zero net force acting on it.
There is no single inertial frame of reference that is preferred over all others
for formulating Newton’s laws. If one frame is inertial, then every other frame
moving relative to it with constant velocity is also inertial. Viewed in this light,
the state of rest and the state of motion with constant velocity are not very different;
both occur when the vector sum of forces acting on the body is zero.
TesT your undersTanding oF secTion 4.2 In which of the following situations is there zero net force on the body? (i) An airplane flying due north at a steady
120 m>s and at a constant altitude; (ii) a car driving straight up a hill with a 3° slope at a
constant 90 km>h; (iii) a hawk circling at a constant 20 km>h at a constant height of 15 m
above an open field; (iv) a box with slick, frictionless surfaces in the back of a truck as
the truck accelerates forward on a level road at 5 m>s2. ❙
4.3 newTon’s second Law
Newton’s first law tells us that when a body is acted on by zero net force, the
body moves with constant velocity and zero acceleration. In Fig. 4.12a, a hockey
puck is sliding to the right on wet ice. There is negligible friction, so there are
no horizontal forces acting on the puck; the downward force of gravity and
theSupward normal force exerted by the ice surface sum to zero. So the net force
g F acting on the puck is zero, the puck has zero acceleration, and its velocity is
constant.
109
4.3 Newton’s Second Law
(a) If there is zero net force on the puck, so gF = 0, ...
S
S
v
4.12 Using a hockey puck on a friction-
S
S
S
less surface to explore the
relationship
S
between the net force gF on a body and
S
the resulting acceleration a of the body.
S
v
v
v
S
S
... the puck has zero acceleration (a = 0) and its velocity v is constant.
v
(b) If a constant net force gF acts on the puck in the direction of its motion ...
S
S
S
ΣF
S
ΣF
S
S
ΣF
S
a
S
a
S
v
ΣF
S
a
S
S
ΣF
S
a
a
S
S
S
v
v
v
v
S
... the puck has a constant acceleration a in the same direction as the net force.
(c) If a constant net force gF acts on the puck opposite to the direction of its motion ...
S
S
S
ΣF
S
S
a
ΣF
S
a
S
S
ΣF
S
a
S
S
ΣF
S
a
S
S
ΣF
a
S
v
v
v
v
S
... the puck has a constant acceleration a in the same direction as the net force.
S
v
4.13 A top view of a hockey puck in
uniform circular motion on a frictionless
horizontal surface.
Puck moves at constant speed
around circle.
S
v
S
S
ΣF
S
v
But what happens when the net force is not zero? In Fig. 4.12b we apply a
constant horizontal Sforce to a sliding puck in the same direction that the puck
S
is moving. Then g F is constant and in the same horizontal direction as v. We
find that during the time the force is acting, the velocity of the puck changes at a
constant rate; that is, the puck moves with constant acceleration. The speed
of the
S
S
S
puck increases, so the acceleration a is in the same direction as v and g F.S
In Fig. 4.12c we reverse the direction of the force on the puck so that g F acts
S
opposite to v. In this case as well, the puck has an acceleration; the puck moves
S
more and more slowly to the
right. The acceleration a in this case is to the left,
S
in the same direction as g F.SAs in the previous case, experiment shows that the
acceleration is constant if g F is constant.
We conclude that a net force acting on a body causes the body to accelerate
in the same direction as the net force. If the magnitude of the net force is constant, as in Figs. 4.12b and 4.12c, then so is the magnitude of the acceleration.
These conclusions about net force and acceleration also apply to a body moving along a curved path. For example, Fig. 4.13 shows a hockey puck moving in
a horizontal circle on an ice surface of negligible friction. A rope is attached to
the puck and to a stick in the ice, and this rope exerts an inward tension force of
constant magnitude on the puck. The net force and acceleration are both constant
in magnitude and directed toward the center of the circle. The speed of the puck
is constant, so this is uniform circular motion, as discussed in Section 3.4.
Figure 4.14a shows another experiment to explore the relationship between
acceleration and net force. We apply a constant horizontal force to a puck on a
frictionless horizontal surface, using the spring balance described in Section 4.1
with the spring stretched a constant amount. As in Figs. 4.12b and 4.12c, this horizontal force equals the net force on the puck. If we change the magnitude of the
net force, the acceleration changes in the same proportion. Doubling the net force
doubles the acceleration (Fig. 4.14b), halving the net force halves the acceleration
(Fig. 4.14c), and so on. Many such experiments show that for any given body, the
magnitude of the acceleration is directly proportional to the magnitude of the
net force acting on the body.
a
S
S
a
ΣF
Rope
S
ΣF
S
a
S
v
S
At all points,
the
acceleration
a and the net
S
force ΣF point in the same direction—always
toward the center of the circle.
4.14 The magnitude of a body’s acceleraS
tion a is directly proportional
to the magS
nitude of the net force gF acting on the
body of mass m.
S
(a) A constant net force ΣF causes a
S
constant acceleration a.
S
a
m
x
S
S
ΣF = F1
(b) Doubling the net force doubles the
acceleration.
S
2a
m
S
S
ΣF = 2F1
x
(c) Halving the force halves the
acceleration.
1S
2a
m
S
S
ΣF = 12 F1
x
110
Chapter 4 Newton’s Laws of Motion
Mass and Force
Our results mean that for a given body, the ratio of the magnitude 0 g F 0 of the
S
net force to the magnitude a = 0 a 0 of the acceleration is constant, regardless of
the magnitude of the net force. We call this ratio the inertial mass, or simply the
mass, of the body and denote it by m. That is,
0 gF 0
S
S
m =
a
or
0 g F 0 = ma
S
0 gF 0
S
or
a =
m
(4.4)
Mass is a quantitative measure of inertia, which we discussed in Section 4.2.
The last of the equations in Eqs. (4.4) says that the greater a body’s mass, the
more the body “resists” being accelerated. When you hold a piece of fruit in your
hand at the supermarket and move it slightly up and down to estimate its heft,
you’re applying a force and seeing how much the fruit accelerates up and down
in response. If a force causes a large acceleration, the fruit has a small mass; if
the same force causes only a small acceleration, the fruit has a large mass. In the
same way, if you hit a table-tennis ball and then a basketball with the same force,
the basketball has much smaller acceleration because it has much greater mass.
The SI unit of mass is the kilogram. We mentioned in Section 1.3 that the
kilogram is officially defined to be the mass of a cylinder of platinum–iridium
alloy kept in a vault near Paris (Fig. 1.4). We can use this standard kilogram,
along with Eqs. (4.4), to define the newton:
One newton is the amount of net force that gives an acceleration of 1 meter per
second squared to a body with a mass of 1 kilogram.
4.15 For a given net force gF acting on a
S
body, the acceleration is inversely proportional to the mass of the body. Masses add
like ordinary scalars.
S
(a) A known force ΣF causes an object
S
with mass m1 to have an acceleration a1.
S
a1
S
ΣF
x
m1
S
(b) Applying the same force ΣF to a
second object and noting the acceleration
allow us to measure the mass.
S
a2
S
ΣF
m2
x
(c) When the two objects are fastened
together, the same method shows that
their composite mass is the sum of their
individual masses.
S
a3
S
ΣF
m1 + m2
x
This definition allows us to calibrate the spring balances and other instruments
used to measure forces. Because of the way we have defined the newton, it is
related to the units of mass, length, and time. For Eqs. (4.4) to be dimensionally
consistent, it must be true that
or
1 newton = 11 kilogram211 meter per second squared2
1 N = 1 kg # m>s2
We will use this relationship many times in the next few chapters, so keep it in mind.
We can also use Eqs. (4.4) to compare a mass with the standard
mass and thus
S
to measure masses. Suppose we apply a constant net force g F to a body having
a known mass m 1 and we find an acceleration of magnitude a1 (Fig. 4.15a). We
then apply the same force to another body having an unknown mass m 2 , and we
find an acceleration of magnitude a2 (Fig. 4.15b). Then, according to Eqs. (4.4),
m 1 a1 = m 2 a2
m2
a1
=
a2
m1
(same net force)
(4.5)
For the same net force, the ratio of the masses of two bodies is the inverse of the
ratio of their accelerations. In principle we could use Eq. (4.5) to measure an
unknown mass m 2 , but it is usually easier to determine mass indirectly by measuring the body’s weight. We’ll return to this point in Section 4.4.
When two bodies with masses m 1 and m 2 are fastened together, we find that
the mass of the composite body is always m 1 + m 2 (Fig. 4.15c). This additive
property of mass may seem obvious, but it has to be verified experimentally.
Ultimately, the mass of a body is related to the number of protons, electrons, and
neutrons it contains. This wouldn’t be a good way to define mass because there
is no practical way to count these particles. But the concept of mass is the most
fundamental way to characterize the quantity of matter in a body.
4.3 Newton’s Second Law
111
Stating Newton’s Second Law
Experiment shows that the net force Son Sa body
is what causes that body to acS
celerate. If a combination of forces F1 , F2 , F3 , and so on is applied to a body,
S
the body will have the same acceleration vector a as when
only
a single
force is
S
S
S
applied, if that single force is equal to the vector sum F1 + F2 + F3 + P . In
other words, the principle of superposition of forces (see Fig. 4.4) also holds true
when the net force is not zero and the body is accelerating.
Equations (4.4) relate the magnitude of the net force on a body to the magnitude of the acceleration that it produces. We have also seen that the direction of
the net force is the same as the direction of the acceleration, whether the body’s
path is straight or curved. What’s more, the forces that affect a body’s motion are
external forces, those exerted on the body by other bodies in its environment.
Newton wrapped up all these results into a single concise statement that we now
call Newton’s second law of motion:
NewtoN’S SecoNd Law of motioN: If a net external force acts on a
body, the body accelerates. The direction of acceleration is the same as the
direction of the net force. The mass of the body times the acceleration vector of
the body equals the net force vector.
In symbols,
Newton’s second law:
If there is a net force on a body ...
g F = ma
S
... the body accelerates in same
direction as the net force.
Mass of body
S
(4.6)
An alternative statement is that the acceleration of a body is equal to the net
force acting on the body divided by the body’s mass:
gF
m
S
S
a =
Newton’s second law is a fundamental law of nature, the basic relationship between force and motion. Most of the remainder of this chapter and all of the next
are devoted to learning how to apply this principle in various situations.
Equation (4.6) has many practical applications (Fig. 4.16). You’ve actually
been using it all your life to measure your body’s acceleration. In your inner ear,
microscopic hair cells sense the magnitude and direction of the force that they
must exert to cause small membranes to accelerate along with the rest of your
body. By Newton’s second law, the acceleration of the membranes—and hence
that of your body as a whole—is proportional to this force and has the same direction. In this way, you can sense the magnitude and direction of your acceleration even with your eyes closed!
Using Newton’s Second Law
There are at least four aspects of Newton’s second law that deserve special attention. First, Eq. (4.6) is a vector equation. Usually we will use it in component
form, with a separate equation for each component of force and the corresponding component of acceleration:
Newton’s second law:
Each component of net force on a body ...
gFx = max
gFy = may
gFz = maz
(4.7)
... equals body’s mass times the corresponding acceleration component.
This set of component equations is equivalent to the single vector Eq. (4.6).
Second, the statement of Newton’s second law refers to external forces. It’s
impossible for a body to affect its own motion by exerting a force on itself; if it
were possible, you could lift yourself to the ceiling by pulling
up on your belt!
S
That’s why only external forces are included in the sum g F in Eqs. (4.6) and (4.7).
4.16 The design of high-performance
motorcycles depends fundamentally on
Newton’s second law. To maximize the
forward acceleration, the designer makes
the motorcycle as light as possible (that is,
minimizes the mass) and uses the most
powerful engine possible (thus maximizing the forward force).
Lightweight
body (small m)
Powerful engine (large F)
application Blame Newton’s
Second Law This car stopped because of
Newton’s second law: The tree exerted an
external force on the car, giving the car an
acceleration that changed its velocity to zero.
112
Chapter 4 Newton’s Laws of Motion
Demo
Third, Eqs. (4.6) and (4.7) are valid only when the mass m is constant. It’s
easy to think of systems whose masses change, such as a leaking tank truck, a
rocket ship, or a moving railroad car being loaded with coal. Such systems are
better handled by using the concept of momentum; we’ll get to that in Chapter 8.
Finally, Newton’s second law is valid in inertial frames of reference only, just
like the first law. Thus it is not valid in the reference frame of any of the accelerating vehicles in Fig. 4.10; relative to any of these frames, the passenger
accelerates even though the net force on the passenger is zero. We will usually
assume that the earth is an adequate approximation to an inertial frame, although
because of its rotation and orbital motion it is not precisely inertial.
cauTion ma is not a force Even though the vector ma is equal to the vector sum gF
S
S
S
S
of all the forces acting on the body, the vector ma is not a force. Acceleration is a result
of a nonzero net force; it is not a force itself. It’s “common sense” to think that there is
a “force of acceleration” that pushes you back into your seat when your car accelerates
forward from rest. But there is no such force; instead, your inertia causes you to tend to
stay at rest relative to the earth, and the car accelerates around you (see Fig. 4.10a). The
“common sense” confusion arises from trying to apply Newton’s second law where it isn’t
valid—in the noninertial reference frame of an accelerating car. We will always examine
motion relative to inertial frames of reference only. ❙
In learning how to use Newton’s second law, we will begin in this chapter
with examples of straight-line motion. Then in Chapter 5 we will consider more
general cases and develop more detailed problem-solving strategies.
Solution
ExamplE 4.4 DEtErmining accElEration from forcE
A worker applies a constant horizontal force with magnitude 20 N
to a box with mass 40 kg resting on a level floor with negligible
friction. What is the acceleration of the box?
soLuTion
idenTiFy and seT up: This problem involves force and acceleration, so we’ll use Newton’s second law. In any problem involving forces, the first steps are to choose a coordinate system and
to identify all of the forces acting on the body in question. It’s
usually convenient to take one axis either along or opposite the
direction of the body’s acceleration, which in this case is horizontal. Hence we take the + x-axis to be in the direction of the applied
horizontal force (which is the direction in which the box accelerates) and the + y-axis to be upward (Fig. 4.17). In most force problems that you’ll encounter (including this one), the force vectors
all lie in a plane, so the z-axis isn’t used.
S
The forces acting on the box are (i) the horizontal force F
S
exerted by the worker, of magnitude 20 N; (ii) the weight w of
the box—that is, the downward gravitational force exerted by the
S
earth; and (iii) the upward supporting force n exerted by the floor.
S
As in Section 4.2, we call n a normal force because it is normal
(perpendicular) to the surface of contact. (We use an italic letter n
to avoid confusion with the abbreviation N for newton.) Friction is
negligible, so no friction force is present.
The box doesn’t move vertically, so the y-acceleration is zero:
ay = 0. Our target variable is the x-acceleration, ax . We’ll find it
by using Newton’s second law in component form, Eqs. (4.7).
execuTe: From Fig. 4.17 only the 20-N force exerted by the
worker has a nonzero x-component. Hence the first of Eqs. (4.7)
tells us that
gFx = F = 20 N = max
The x-component of acceleration is therefore
4.17 Our sketch for this problem. The tiles under the box are
freshly waxed, so we assume that friction is negligible.
The box has no vertical acceleration, so the vertical
components of the net force sum to zero. Nevertheless, for
completeness, we show the vertical forces acting on the box.
ax =
20 kg # m>s2
gFx
20 N
= 0.50 m>s2
=
=
m
40 kg
40 kg
evaLuaTe: The acceleration is in the +x-direction, the same direction as the net force. The net force is constant, so the acceleration is also constant. If we know the initial position and velocity of
the box, we can find its position and velocity at any later time from
the constant-acceleration equations of Chapter 2.
To determine ax, we didn’t need the y-component of Newton’s
second law from Eqs. (4.7), gFy = may . Can you use this equation to show that the magnitude n of the normal force in this situation is equal to the weight of the box?
4.3 Newton’s Second Law
113
Solution
ExamplE 4.5 DEtErmining forcE from accElEration
A waitress shoves a ketchup bottle with mass 0.45 kg to her right
along a smooth, level lunch counter. The bottle leaves her hand
moving at 2.8 m>s, then slows down as it slides because of a constant horizontal friction force exerted on it by the countertop. It
slides for 1.0 m before coming to rest. What are the magnitude
and direction of the friction force acting on the bottle?
SoLUtioN
ideNtifY and Set Up: This problem involves forces and accel-
eration (the slowing of the ketchup bottle), so we’ll use Newton’s
second law to solve it. As in Example 4.4, we choose a coordinate
system and identify the forces acting on the bottle (Fig. 4.18). We
choose the + x-axis to be in the direction that the bottle slides, and
take the origin to Sbe where the bottle leaves the waitress’s hand.
The friction force f slows the bottle down, so its direction must be
opposite the direction of the bottle’s velocity (see Fig. 4.12c).
Our target variable is the magnitude f of the friction force. We’ll
find it by using the x-component of Newton’s second law from
4.18 Our sketch for this problem.
We draw one diagram for the bottle’s motion and one showing the forces
on the bottle.
Eqs. (4.7). We aren’t told the x-component of the bottle’s acceleration, ax , but we know that it’s constant because the friction force that
causes the acceleration is constant. Hence we can use a constantacceleration formula from Section 2.4 to calculate ax . We know the
bottle’s initial and final x-coordinates 1x0 = 0 and x = 1.0 m2 and
its initial and final x-velocity 1v0x = 2.8 m>s and vx = 02, so the
easiest equation to use is Eq. (2.13), v x 2 = v 0x2 + 2ax1x - x02.
execUte: We solve Eq. (2.13) for ax:
ax =
10 m>s22 - 12.8 m>s22
v x 2 - v 0x 2
=
= - 3.9 m>s2
21x - x02
211.0 m - 0 m2
The negative sign means that the bottle’s acceleration is toward the
left in Fig. 4.18, opposite to its velocity; this is as it must be, because the bottle is slowing down. The net force in the x-direction
is the x-component -f of the friction force, so
gFx = -f = max = 10.45 kg21- 3.9 m>s22
= -1.8 kg # m>s2 = - 1.8 N
The negative sign shows that the net force on the bottle is toward
the left. The magnitude of the friction force is f = 1.8 N.
evaLUate: As a check on the result, try repeating the calculation
with the +x-axis to the left in Fig. 4.18. You’ll find that gFx is
equal to +f = + 1.8 N (because the friction force is now in the
+x-direction), and again you’ll find f = 1.8 N. The answers for
the magnitudes of forces don’t depend on the choice of coordinate
axes!
Some Notes on Units
A few words about units are in order. In the cgs metric system (not used in this
book), the unit of mass is the gram, equal to 10-3 kg, and the unit of distance is
the centimeter, equal to 10-2 m. The cgs unit of force is called the dyne:
4.19 Despite its name, the English unit of
mass has nothing to do with the type of
slug shown here. A common garden slug
has a mass of about 15 grams, or about
10-3 slug.
1 dyne = 1 g # cm>s2 = 10-5 N
In the British system, the unit of force is the pound (or pound-force) and the
unit of mass is the slug (Fig. 4.19). The unit of acceleration is 1 foot per second
squared, so
1 pound = 1 slug # ft>s2
The official definition of the pound is
1 pound = 4.448221615260 newtons
It is handy to remember that a pound is about 4.4 N and a newton is about
0.22 pound. Another useful fact: A body with a mass of 1 kg has a weight of
about 2.2 lb at the earth’s surface.
Table 4.2 lists the units of force, mass, and acceleration in the three systems.
teSt YoUr UNderStaNdiNg of SectioN 4.3 Rank the following situations
in order of the magnitude of the object’s acceleration, from lowest to highest. Are there
any cases that have the same magnitude of acceleration? (i) A 2.0-kg object acted on by a
2.0-N net force; (ii) a 2.0-kg object acted on by an 8.0-N net force; (iii) an 8.0-kg object
acted on by a 2.0-N net force; (iv) an 8.0-kg object acted on by a 8.0-N net force. ❙
Units of force, mass,
Table 4.2 and acceleration
System
of Units
SI
cgs
British
Force
Mass
Acceleration
newton
(N)
dyne
(dyn)
pound
(lb)
kilogram
(kg)
gram
(g)
slug
m>s2
cm>s2
ft>s2
114
Chapter 4 Newton’s Laws of Motion
4.4 MaSS and Weight
4.20 Relating the mass and weight of
a body.
Falling body,
mass m
Hanging body,
mass m
S
T
S
S
S
a=0
a=g
Weight
S
S
w = mg
S
S
ΣF = w
Weight
S
S
w = mg
S
ΣF = 0
S
S
• The relationship of mass to weight: w = mg.
• This relationship is the same whether a body
is falling or stationary.
One of the most familiar forces is the weight of a body, which is the gravitational
force that the earth exerts on the body. (If you are on another planet, your weight
is the gravitational force that planet exerts on you.) Unfortunately, the terms mass
and weight are often misused and interchanged in everyday conversation. It is
absolutely essential for you to understand clearly the distinctions between these
two physical quantities.
Mass characterizes the inertial properties of a body. Mass is what keeps the
china on the table when you yank the tablecloth out from under it. The greater
the mass, the greater the force needed
to cause a given acceleration; this is reS
S
flected in Newton’s second law, g F = m a.
Weight, on the other hand, is a force exerted on a body by the pull of the
earth. Mass and weight are related: Bodies that have large mass also have large
weight. A large stone is hard to throw because of its large mass, and hard to lift
off the ground because of its large weight.
To understand the relationship between mass and weight, note that a freely
falling body has an acceleration of magnitude g (see Section 2.5). Newton’s second law tells us that a force must act to produce this acceleration. If a 1-kg body
falls with an acceleration of 9.8 m>s2, the required force has magnitude
F = ma = 11 kg219.8 m>s22 = 9.8 kg # m>s2 = 9.8 N
The force that makes the body accelerate downward is its weight. Any body
near the surface of the earth that has a mass of 1 kg must have a weight of 9.8 N
to give it the acceleration we observe when it is in free fall. More generally,
Magnitude of
weight of a body
caution a body’s weight acts at all
times The weight of a body acts on the
body all the time, whether it is in free fall
or not. If we suspend an object from a rope,
it is in equilibrium and its acceleration is
zero. But its weight, given by Eq. (4.9), is
still pulling down on it (Fig. 4.20). In this
case the rope pulls up on the object, applying an upward force. The vector sum of the
forces is zero, but the weight still acts. ❙
Magnitude of acceleration
due to gravity
(4.8)
Hence the magnitude w of a body’s weight is directly proportional to its mass m.
The weight of a body is a force, a vector quantity, and we can write Eq. (4.8) as a
vector equation (Fig. 4.20):
S
S
w = mg
(4.9)
S
Remember that g is the magnitude of g, the acceleration due to gravity, so g is
always a positive number, by definition. Thus w, given by Eq. (4.8), is the magnitude
of the weight and is also always positive.
Solution
concEptual ExamplE 4.6
Mass of body
w = mg
nEt forcE and accElEration in frEE fall
In Example 2.6 of Section 2.5, a one-euro coin was dropped from
rest from the Leaning Tower of Pisa. If the coin falls freely, so that
the effects of the air are negligible, how does the net force on the
coin vary as it falls?
Solution
In free fall, the acceleration a of the coin is constant
and equal to
S
S
S
g. Hence by Newton’s second law the net force gF = ma is also
S
S
constant and equal to mg, which is the coin’s weight w (Fig. 4.21).
The coin’s velocity changes as it falls, but the net force acting on it
is constant. (If this surprises you, reread Conceptual Example 4.3.)
The net force on a freely falling coin is constant even if you
initially toss it upward. The force that your hand exerts on the coin
to toss it is a contact force, and it disappears the instant the coin
leaves your hand. From then on, the only force acting on the coin
S
is its weight w.
4.21 The acceleration of a freely falling object is constant, and
so is the net force acting on the object.
S
S
S
a=g
S
S
ΣF = w
4.4 Mass and Weight
variation of g with Location
2
We will use g = 9.80 m>s for problems set on the earth (or, if the other data
in the problem are given to only two significant figures, g = 9.8 m>s2). In fact,
the value of g varies somewhat from point to point on the earth’s surface—from
about 9.78 to 9.82 m>s2—because the earth is not perfectly spherical and because
of effects due to its rotation and orbital motion. At a point where g = 9.80 m>s2,
the weight of a standard kilogram is w = 9.80 N. At a different point, where
g = 9.78 m>s2, the weight is w = 9.78 N but the mass is still 1 kg. The weight of
a body varies from one location to another; the mass does not.
If we take a standard kilogram to the surface of the moon, where the acceleration of free fall (equal to the value of g at the moon’s surface) is 1.62 m>s2,
its weight is 1.62 N but its mass is still 1 kg (Fig. 4.22). An 80.0-kg astronaut
has a weight on earth of 180.0 kg219.80 m>s22 = 784 N, but on the moon the
astronaut’s weight would be only 180.0 kg211.62 m>s22 = 130 N. In Chapter 13
we’ll see how to calculate the value of g at the surface of the moon or on other
worlds.
4.22 The weight of a 1-kilogram mass
(a) on earth and (b) on the moon.
(a)
20 0 2
18
4
16
6
14
8
12 10
m = 1.00 kg
On earth:
g = 9.80 m>s2
w = mg = 9.80 N
(b)
20 0 2
18
4
16
6
14
8
12 10
Measuring Mass and weight
In Section 4.3 we described a way to compare masses by comparing their accelerations when they are subjected to the same net force. Usually, however, the
easiest way to measure the mass of a body is to measure its weight, often by comparing with a standard. Equation (4.8) says that two bodies that have the same
weight at a particular location also have the same mass. We can compare weights
very precisely; the familiar equal-arm balance (Fig. 4.23) can determine with
great precision (up to 1 part in 106) when the weights of two bodies are equal and
hence when their masses are equal.
The concept of mass plays two rather different roles in mechanics. The weight
of a body (the gravitational force acting on it) is proportional to its mass; we
call the property related to gravitational interactions gravitational mass. On the
other hand, we call the inertial property that appears in Newton’s second law the
inertial mass. If these two quantities were different, the acceleration due to gravity might well be different for different bodies. However, extraordinarily precise
experiments have established that in fact the two are the same to a precision of
better than one part in 1012.
115
On the moon:
g = 1.62 m>s2
w = mg = 1.62 N
m = 1.00 kg
4.23 An equal-arm balance determines
the mass of a body (such as an apple) by
comparing its weight to a known weight.
d
wunknown
d
wknown
cauTion Don’t confuse mass and weight The SI units for mass and weight are often
misused in everyday life. For example, it’s incorrect to say “This box weighs 6 kg”; what
this really means is that the mass of the box, probably determined indirectly by weighing,
is 6 kg. Avoid this sloppy usage in your own work! In SI units, weight (a force) is measured
in newtons, while mass is measured in kilograms. ❙
Solution
ExamplE 4.7 maSS anD wEight
A 2.49 * 104 N Rolls-Royce Phantom traveling in the +x-direction
makes an emergency stop; the x-component of the net force acting
on it is - 1.83 * 104 N. What is its acceleration?
soLuTion
idenTiFy and seT up: Our target variable is the x-component
of the car’s acceleration, ax . We use the x-component portion of
Newton’s second law, Eqs. (4.7), to relate force and acceleration.
To do this, we need to know the car’s mass. The newton is a unit
for force, however, so 2.49 * 104 N is the car’s weight, not its
mass. Hence we’ll first use Eq. (4.8) to determine the car’s mass
from its weight. The car has a positive x-velocity and is slowing
down, so its x-acceleration will be negative.
execuTe: The mass of the car is
m =
2.49 * 104 kg # m>s2
w
2.49 * 104 N
=
=
g
9.80 m>s2
9.80 m>s2
= 2540 kg
Continued
116
Chapter 4 Newton’s Laws of Motion
Then g Fx = max gives
ax =
- 1.83 * 104 kg # m>s2
gFx
- 1.83 * 104 N
=
=
m
2540 kg
2540 kg
= - 7.20 m>s2
evaLuaTe: The negative sign means that the acceleration vector
points in the negative x-direction, as we expected. The magnitude
of this acceleration is pretty high; passengers in this car will experience a lot of rearward force from their shoulder belts.
This acceleration equals -0.735g. The number - 0.735 is also
the ratio of -1.83 * 104 N (the x-component of the net force)
to 2.49 * 104 N (the weight). In fact, the acceleration of a body,
expressed as a multiple of g, is always equal to the ratio of the net
force on the body to its weight. Can you see why?
TesT your undersTanding oF secTion 4.4 Suppose an astronaut landed
on a planet where g = 19.6 m>s2. Compared to earth, would it be easier, harder, or just
as easy for her to walk around? Would it be easier, harder, or just as easy for her to catch
a ball that is moving horizontally at 12 m>s? (Assume that the astronaut’s spacesuit is a
lightweight model that doesn’t impede her movements in any way.) ❙
4.5 newTon’s Third Law
4.24 Newton’s third law of motion.
S
If body A exerts force FA on B on body B
(for example, a foot kicks a ball) ...
B
S
FA on B
A force acting on a body is always the result of its interaction with another body,
so forces always come in pairs. You can’t pull on a doorknob without the doorknob pulling back on you. When you kick a football, the forward force that your
foot exerts on the ball launches it into its trajectory, but you also feel the force the
ball exerts back on your foot.
In each of these cases, the force that you exert on the other body is in the
opposite direction to the force that body exerts on you. Experiments show that
whenever two bodies interact, the two forces that they exert on each other are
always equal in magnitude and opposite in direction. This fact is called Newton’s
third law of motion:
newTon’s Third Law oF MoTion: If body A exerts a force on body B (an
“action”), then body B exerts a force on body A (a “reaction”). These two forces
have the same magnitude but are opposite in direction. These two forces act on
different bodies.
S
A
S
FB on A
... then body SB necessarily
exerts force FB on A on body A
(ball kicks back on foot).
The two forces have same
magnitudeS
S
but opposite directions: FA on B = −FB on A.
For example, in Fig. 4.24 FA on B is theS force applied by body A (first subscript) on body B (second subscript), and FB on A is the force applied by body B
(first subscript) on body A (second subscript). In equation form,
Newton’s third law:
When two bodies
A and B exert forces
on each other ...
S
S
FA on B = −FB on A
... the two forces have
same magnitude but
opposite directions.
(4.10)
Note: The two forces act on different bodies.
?
Demo
It doesn’t matter whether one body is inanimate (like the soccer ball in
Fig. 4.24) and the other is not (like the kicker’s foot): They necessarily exert
forces on each other that obey Eq. (4.10).
In the statement of Newton’s
third law,
“action” and “reaction” are the two
S
S
opposite forces (in Fig. 4.24, F A on B and FB on A); we sometimes refer to them as an
action–reaction pair. This is not meant to imply any cause-and-effect relationship; we can consider either force as the “action” and the other as the “reaction.”
We often say simply that the forces are “equal and opposite,” meaning that they
have equal magnitudes and opposite directions.
cauTion the two forces in an action–reaction pair act on different bodies We stress that
the two forces described in Newton’s third law act on different bodies. This is important
in problems involving Newton’s first or second law, which involve the forces that act on a
single body. For instance, the net force
on the soccer ball in Fig. 4.24 is the vector sum of
S
the weight
of
the
ball
and
the
force
F
A
on
B exerted by the kicker. You wouldn’t include the
S
force FB on A because this force acts on the kicker, not on the ball. ❙
4.5 Newton’s third Law
117
In Fig. 4.24 the action and reaction forces are contact forces that are present only when the two bodies are touching. But Newton’s third law also applies
to long-range forces that do not require physical contact, such as the force of
gravitational attraction. A table-tennis ball exerts an upward gravitational force
on the earth that’s equal in magnitude to the downward gravitational force the
earth exerts on the ball. When you drop the ball, both the ball and the earth accelerate toward each other. The net force on each body has the same magnitude,
but the earth’s acceleration is microscopically small because its mass is so great.
Nevertheless, it does move!
Solution
Which forcE iS grEatEr?
concEptual ExamplE 4.8
After your sports car breaks down, you start to push it to the nearest repair shop. While the car is starting to move, how does the
force you exert on the car compare to the force the car exerts on
you? How do these forces compare when you are pushing the car
along at a constant speed?
Solution
Newton’s third law says that in both cases, the force you exert on
the car is equal in magnitude and opposite in direction to the force
the car exerts on you. It’s true that you have to push harder to get the
car going than to keep it going. But no matter how hard you push
on the car, the car pushes just as hard back on you. Newton’s third
law gives the same result whether the two bodies are at rest, moving with constant velocity, or accelerating.
You may wonder how the car “knows” to push back on you
with the same magnitude of force that you exert on it. It may help
to visualize the forces you and the car exert on each other as interactions between the atoms at the surface of your hand and the
atoms at the surface of the car. These interactions are analogous
to miniature springs between adjacent atoms, and a compressed
spring exerts equally strong forces on both of its ends.
Fundamentally, though, the reason we know that objects of
different masses exert equally strong forces on each other is that
experiment tells us so. Physics isn’t merely a collection of rules
and equations; rather, it’s a systematic description of the natural
world based on experiment and observation.
Solution
applying nEWton’S third laW: objEctS at rESt
concEptual ExamplE 4.9
An apple sits at rest on a table, in equilibrium. What forces act on
the apple? What is the reaction force to each of the forces acting
on the apple? What are the action–reaction pairs?
Solution
Figure 4.25b shows one of the action–reaction pairs involving
the apple. As the earth pulls down on the apple, with force
S
Fearth on
, the apple exerts an equally strong upward pull on the
S apple
earth Fapple on earth. By Newton’s third law (Eq. 4.10) we have
S
S
Figure 4.25a shows the forces acting on the apple. Fearth on apple is
the weight of the apple—that is, the downward
gravitational force
S
exerted by the earth on the apple. Similarly, Ftable on apple is the upward normal force exerted by the table on the apple.
S
Fapple on earth = −Fearth on apple
S
Also, as the table pushes up on the apple with force Ftable on apple ,
S
the corresponding reaction is the downward force Fapple on table
4.25 The two forces in an action–reaction pair always act on different bodies.
(a) The forces acting on the apple
(b) The action–reaction pair for
the interaction between the
apple and the earth
(c) The action–reaction pair for
the interaction between the apple
and the table
(d) We eliminate one of the forces
acting on the apple.
S
S
Ftable on apple
S
Ftable on apple
Ftable on apple = 0
S
S
S
Fearth on apple
Fearth on apple
S
Fapple on table
Fearth on apple
S
Fapple on earth
S
Fapple on earth
S
S
Fapple on earth = −Fearth on apple
Table
removed
S
S
Fapple on table = −Ftable on apple
Action–reaction pairs always represent a
mutual interaction of two different objects.
The two forces on the apple cannot
be an action–reaction pair, because
they act on the same object.
Continued
118
Chapter 4 Newton’s Laws of Motion
exerted by the apple on the table (Fig. 4.25c). For this action–
reaction pair we have
S
S
Fapple on table = −Ftable on apple
S
The two forces acting on the apple, Ftable on apple and
Fearth on apple , are not an action–reaction pair, despite being equal
in magnitude and opposite in direction. They do not represent the
mutual interaction of two bodies; they are two different forces
S
applying nEwton’S third law: objEctS in motion
A stonemason drags a marble block across a floor by pulling on a
rope attached to the block (Fig. 4.26a). The block is not necessarily in equilibrium. How are the various forces related? What are
the action–reaction pairs?
Solution
We’ll use the subscripts B for the block,
R for the rope, and M
S
for the mason. In Fig. 4.26b the vector FM on R represents the force
exerted by the
mason on the rope. The corresponding reaction
S
is
the
force
F
R
on
M exerted by the rope on the mason. Similarly,
S
FR on B represents the force exerted by Sthe rope on the block, and
the corresponding reaction is the force FB on R exerted by the block
on the rope. The forces in each action–reaction pair are equal and
opposite:
S
S
FR on M = −FM on R
S
and
S
S
FB on R = −FR on B
S
Forces FM on R and FB on R (Fig. 4.26c) are not an action–
reaction pair, because both of these forces act on the same body
(the rope); an action and its reaction
must always
act on different
S
S
bodies. Furthermore, the forces FM on R and FB on R are not necessarily equal in magnitude. Applying Newton’s second law to the
rope, we get
gF = FM on R + FB on R = m rope arope
S
S
Solution
concEptual ExamplE 4.10
acting on the same body. Figure 4.25d shows another way to see
this. If Swe suddenly yankS the table out from under the apple, the
forces Fapple on table and Ftable on apple suddenly become zero, but
S
S
F apple on earth and Fearth on apple are unchanged (the gravitational
S
interaction is still present). Because
Ftable on apple is now zero, it can’t
S
be the negative of the nonzero Fearth on apple, and these two forces
can’t be an action–reaction pair. The two forces in an action–
reaction pair never act on the same body.
S
S
differentS magnitudeSthan FB on R. By contrast, the action–reaction
forces
FM on SR and FR on M are always equal in magnitude, as are
S
FR on B and FB on R. Newton’s third law holds whether or not the
bodies are accelerating.
In the
special case
in which the rope is in equilibrium, the
S
S
forces FM on R and FB on R are equal in magnitude, and they are
opposite in direction. But this is an example of Newton’s first
law, not his third; these are two forces on the same body, not
forces of two bodies on each other. Another way to look at this
S
is
that in equilibrium,
arope = 0 in the previous equation. Then
S
S
FB on R = −FM on R because of Newton’s first or second law.
Another special case is if the rope is accelerating but has negligibly small mass compared to that of the block or the mason.
In
this case,S m rope = 0 in the previous equation, so Sagain
S
FB on R = −FM on R. Since Newton’s third law says that FB on R
S
always equals −FR on B (they are an action–reaction pair), in this
S
S
“massless-rope” case FR on B also equals FM on R.
For both the “massless-rope” case and the case of the rope
in equilibrium, the force of the rope on the block is equal in
magnitude and direction to the force of the mason on the rope
(Fig. 4.26d). Hence we can think of the rope as “transmitting” to
the block the force the mason exerts on the rope. This is a useful
point of view, but remember that it is valid only when the rope has
negligibly small mass or is in equilibrium.
S
If the block and rope are accelerating (speeding
up or slowing
S
down), the rope is not in equilibrium, and FM on R must have a
4.26 Identifying the forces that act when a mason pulls on a rope attached to a block.
(a) The block, the rope, and the mason
(b) The action–reaction pairs
S
FR on M
S
S
FB on R FR on B
S
FM on R
(c) Not an action–reaction pair
S
FB on R
S
FM on R
These forces cannot be
an action–reaction pair
because they act on the
same object (the rope).
(d) Not necessarily equal
S
FR on B
S
FM on R
These forces are equal only if
the rope is in equilibrium (or
can be treated as massless).
4.5 Newton’s third Law
Solution
concEptUal ExamplE 4.11
119
a nEwton’S thirD law paraDox?
We saw in Conceptual Example 4.10 that the stonemason pulls
as hard on the rope–block combination as that combination pulls
back on him. Why, then, does the block move while the stonemason remains stationary?
4.27 The horizontal forces acting on the block–rope com-
bination (left) and the mason (right). (The vertical forces are
not shown.)
These forces are an action–reaction
pair. They have the same magnitude
but act on different objects.
soLuTion
To resolve this seeming paradox, keep in mind the difference between Newton’s second and third laws. The only forces involved in
Newton’s second law are those that act on a given body. The vector
sum of these forces determines the body’s acceleration, if any. By
contrast, Newton’s third law relates the forces that two different
bodies exert on each other. The third law alone tells you nothing
about the motion of either body.
If the rope–block combination is initially
at rest, it begins to
S
slide if the stonemason exerts a force FM on R that is greater in
magnitude than the friction force that the floor exerts on the block
(Fig. 4.27). (The block has a smooth underside, which minimizes
friction.) Then there is a net force to the right on the rope–block
combination, and it accelerates to the right. By contrast, the stonemason doesn’t move because the net force acting on him is zero.
His shoes have nonskid soles that don’t slip on the floor, so the
friction force that the floor exerts Son him is strong enough to balance the pull of the rope on him, FR on M. (Both the block and the
stonemason also experience a downward force of gravity and an
upward normal force exerted by the floor. These forces balance
each other, so we haven’t included them in Fig. 4.27.)
Once the block is moving, the stonemason doesn’t need to pull
as hard; he must exert only enough force to balance the friction
force on the block. Then the net force on the moving block is zero,
and by Newton’s first law the block continues to move toward the
mason at a constant velocity.
Friction force
of floor on
block
S
Friction force
of floor on
mason
S
FM on R FR on M
Block + rope
Mason
The
block begins sliding if
S
FM on R overcomes the
friction force on the block.
The
mason remains at rest if
S
FR on M is balanced by the
friction force on the mason.
So the block accelerates but the stonemason doesn’t because
different amounts of friction act on them. If the floor were freshly
waxed, so that there was little friction between the floor and the
stonemason’s shoes, pulling on the rope might start the block sliding to the right and start him sliding to the left.
The moral of this example is that when analyzing the motion of
a body, you must remember that only the forces acting on a body
determine its motion. From this perspective, Newton’s third law is
merely a tool that can help you determine what those forces are.
A body that has pulling forces applied at its ends, such as the rope in Fig. 4.26,
is said to be in tension. The tension at any point along the rope is the magnitude
of the force acting at that point (see Fig. S4.2c). In Fig. S4.26b the tension at the
right end of the rope is the magnitude of SFM on R (or of SFR on M), and the tension
at the left end equals the magnitude of FB on R (or of FR on B). If the rope is in
equilibrium and if no forces act except at its ends, the tension
is the same
at both
S
S
ends and throughout the rope. Thus, if the magnitudes of FB on R and FM on R are
50
N each,S the tension in the rope is 50 N (not 100 N). The total force vector
S
FB on R + FM on R acting on the rope in this case is zero!
We emphasize once again that the two forces in an action–reaction pair never
act on the same body. Remembering this fact can help you avoid confusion about
action–reaction pairs and Newton’s third law.
daTa SpeakS
Force and Motion
When students were given a problem
about forces acting on an object and how
they affect the object’s motion, more than
20% gave an incorrect answer. Common
errors:
●
●
TesT your undersTanding oF secTion 4.5 You are driving a car on a
country road when a mosquito splatters on the windshield. Which has the greater magnitude: the force that the car exerted on the mosquito or the force that the mosquito exerted
on the car? Or are the magnitudes the same? If they are different, how can you reconcile
this fact with Newton’s third law? If they are equal, why is the mosquito splattered while
the car is undamaged? ❙
Confusion about contact forces. If your
fingers push on an object, the force you
exert acts only when your fingers and
the object are in contact. Once contact
is broken, the force is no longer present
even if the object is still moving.
Confusion about Newton’s third law.
The third law relates the forces that two
objects exert on each other. By itself,
this law can’t tell you anything about
two forces that act on the same object.
120
Chapter 4 Newton’s Laws of Motion
4.6 Free-Body diagraMs
Demo
Demo
4.28 The simple act of walking depends
crucially on Newton’s third law. To start
moving forward, you push backward on
the ground with your foot. As a reaction,
the ground pushes forward on your foot
(and hence on your body as a whole) with
a force of the same magnitude. This
external force provided by the ground is
what accelerates your body forward.
cauTion forces in free-body diagrams
For a free-body diagram to be complete,
you must be able to answer this question
for each force: What other body is applying
this force? If you can’t answer that question, you may be dealing with a nonexistent
force. Avoid nonexistent forces such as “the
S
force of acceleration” or “the ma force,”
discussed in Section 4.3. ❙
Newton’s three laws of motion contain all the basic principles we need to solve
a wide variety of problems in mechanics. These laws are very simple in form,
but the process of applying them to specific situations can pose real challenges.
In this brief section we’ll point out three key ideas and techniques to use in any
problems involving Newton’s laws. You’ll learn others in Chapter 5, which also
extends the use of Newton’s laws to cover more complex situations.
Newton’s first Slaw, g F = 0, for an equilibrium situation or Newton’s
S
second law, g F = ma, for a nonequilibrium situation, you must decide at
the beginning to which body you are referring. This decision may sound
trivial, but it isn’t.
S
2. Only forces acting on the body matter. The sum g F includes all the forces
that act on the body in question. Hence, once you’ve chosen the body to
analyze, you have to identify all the forces acting on it. Don’t confuse the
forces acting on a body with the forces exerted by that body on some other
S
body. For example, to analyze a person walking, you would include in g F
the force that the ground exerts on the person as he walks, but not the
force that the person exerts on the ground (Fig. 4.28). These forces form
an action–reaction pair and are related by Newton’s third law, but only
the member
of the pair that acts on the body you’re working with goes
S
into g F.
3. Free-body diagrams are essential to help identify the relevant forces. A
free-body diagram shows the chosen body by itself, “free” of its surroundings, with vectors drawn to show the magnitudes and directions of
all the forces that act on the body. We’ve already shown free-body diagrams in Figs. 4.17, 4.18, 4.20, and 4.25a. Be careful to include all the
forces acting on the body, but be equally careful not to include any forces
that the body exerts on any other body. In particular, the two forces in an
action–reaction pair must never appear in the same free-body diagram because they never act on the same body. Furthermore, never include forces
that a body exerts on itself, since these can’t affect the body’s motion.
1. Newton’s first and second
laws apply to a specific body. Whenever you use
S
When a problem involves more than one body, you have to take the problem apart and draw a separate free-body diagram for each body. For example, Fig. 4.26c shows a separate free-body diagram for the rope in the case in
which the rope is considered massless (so that no gravitational force acts on it).
Figure 4.27 also shows diagrams for the block and the mason, but these are not
complete free-body diagrams because they don’t show all the forces acting on
each body. (We left out the vertical forces—the weight force exerted by the earth
and the upward normal force exerted by the floor.)
In Fig. 4.29 we present three real-life situations and the corresponding complete free-body diagrams. Note that in each situation a person exerts a force on
something in his or her surroundings, but the force that shows up in the person’s
free-body diagram is the surroundings pushing back on the person.
TesT your undersTanding oF secTion 4.6 The buoyancy force shown
in Fig. 4.29c is one half of an action–reaction pair. What force is the other half of this
pair? (i) The weight of the swimmer; (ii) the forward thrust force; (iii) the backward drag
force; (iv) the downward force that the swimmer exerts on the water; (v) the backward
force that the swimmer exerts on the water by kicking. ❙
121
4.6 Free-Body Diagrams
4.29 Examples of free-body diagrams. Each free-body diagram shows all of the external forces that act on the
object in question.
(a)
(b)
S
n
S
To jump up, this
player will push
down against the
floor, increasing
the upward reaction
S
force n of the floor
on him.
S
Fy
Fblock on runner
S
S
w
Fx
S
w
The force of the starting block
on the runner has a vertical
component that counteracts her
weight and a large horizontal
component that accelerates her.
S
w
This player is a
freely falling object.
(c)
S
The water exerts a buoyancy force that
counters the swimmer’s weight.
Fbuoyancy
S
S
Fthrust
Kicking causes the water to
exert a forward reaction force,
or thrust, on the swimmer.
Chapter
Fdrag
S
w
Thrust is countered by drag
forces exerted by the water
on the moving swimmer.
4 Summary
Force as a vector: Force is a quantitative measure
of the interaction between two bodies. It is a vector
quantity. When several forces act on a body, the effect
on its motion is the same as when a single force, equal
to the vector sum (resultant) of the forces, acts on the
body. (See Example 4.1.)
The net force on a body and Newton’s first law:
Newton’s first law states that when the vector sum of
all forces acting on a body (the net force) is zero, the
body is in equilibrium and has zero acceleration. If
the body is initially at rest, it remains at rest; if it is
initially in motion, it continues to move with constant
velocity. This law is valid in inertial frames of reference only. (See Examples 4.2 and 4.3.)
SolutionS to all exampleS
R = gF = F1 + F2 + F3 + P
S
S
S
S
S
R = gF = F1 + F2
(4.1)
S
S
F1
S
S
S
S
F2
gF = 0
S
S
v = constant
(4.3)
S
S
S
F2 = −F1
F1
S
ΣF = 0
122
Chapter 4 Newton’s Laws of Motion
Mass, acceleration, and Newton’s second law: The inertial properties of a body are characterized by its mass.
The acceleration of a body under the action of a given
set of forces is directly proportional to the vector sum
of the forces (the net force) and inversely proportional
to the mass of the body. This relationship is Newton’s
second law. Like Newton’s first law, this law is valid
in inertial frames of reference only. The unit of force
is defined in terms of the units of mass and acceleration. In SI units, the unit of force is the newton (N),
equal to 1 kg # m>s2. (See Examples 4.4 and 4.5.)
S
Weight: The weight w
of a body is the gravitational
gF = ma
S
S
(4.6)
g Fx = max
g Fy = may
(4.7)
w = mg
(4.8)
g Fz = maz
force exerted on it by the earth. Weight is a vector
quantity. The magnitude of the weight of a body at
any specific location is equal to the product of its
mass m and the magnitude of the acceleration due to
gravity g at that location. While the weight of a body
depends on its location, the mass is independent of
location. (See Examples 4.6 and 4.7.)
Newton’s third law and action–reaction pairs:
Newton’s third law states that when two bodies
interact, they exert forces on each other that are equal
in magnitude and opposite in direction. These forces
are called action and reaction forces. Each of these
two forces acts on only one of the two bodies; they
never act on the same body. (See Examples 4.8–4.11.)
S
Mass m
S
S
g
B
(4.10)
S
FA on B
A
S
FB on A
linkS in a chain
A student suspends a chain consisting of three links, each of mass
m = 0.250 kg, from a light rope. The rope is attached to the top
link of the chain, which does not swing. She pulls upward on the
rope, so that the rope applies an upward force of 9.00 N to the chain.
(a) Draw a free-body diagram for the entire chain, considered as a
body, and one for each of the three links. (b) Use the diagrams of
part (a) and Newton’s laws to find (i) the acceleration of the chain,
(ii) the force exerted by the top link on the middle link, and (iii) the
force exerted by the middle link on the bottom link. Treat the rope
as massless.
F1
Solution
bridging problEm
Mass m
w = mg
FA on B = −FB on A
S
S
a = ΣF>m
S
S
S
S
ΣF
S
F2
4. Use your lists to decide how many unknowns there are in this
problem. Which of these are target variables?
execute
5. Write a Newton’s second law equation for each of the four objects, and write a Newton’s third law equation for each action–
reaction pair. You should have at least as many equations as
there are unknowns (see step 4). Do you?
6. Solve the equations for the target variables.
evaluate
Solution guide
identify and Set up
1. There are four objects of interest in this problem: the chain as a
whole and the three individual links. For each of these four
objects, make a list of the external forces that act on it. Besides
the force of gravity, your list should include only forces exerted
by other objects that touch the object in question.
2. Some of the forces in your lists form action–reaction pairs (one
pair is the force of the top link on the middle link and the force
of the middle link on the top link). Identify all such pairs.
3. Use your lists to help you draw a free-body diagram for each of
the four objects. Choose the coordinate axes.
7. You can check your results by substituting them back into the
equations from step 6. This is especially important to do if you
ended up with more equations in step 5 than you used in step 6.
8. Rank the force of the rope on the chain, the force of the top
link on the middle link, and the force of the middle link on the
bottom link in order from smallest to largest magnitude. Does
this ranking make sense? Explain.
9. Repeat the problem for the case in which the upward force that
the rope exerts on the chain is only 7.35 N. Is the ranking in
step 8 the same? Does this make sense?
Discussion Questions
123
Problems
For assigned homework and other learning materials, go to MasteringPhysics®.
., .., ...: Difficulty levels. CP: Cumulative problems incorporating material from earlier chapters. CALC: Problems requiring calculus.
DATA: Problems involving real data, scientific evidence, experimental design, and/or statistical reasoning. bio: Biosciences problems.
discussion QuesTions
Q4.1 Can a body be in equilibrium when only one force acts on
it? Explain.
Q4.2 A ball thrown straight up has zero velocity at its highest
point. Is the ball in equilibrium at this point? Why or why not?
Q4.3 A helium balloon hovers in midair, neither ascending nor
descending. Is it in equilibrium? What forces act on it?
Q4.4 When you fly in an airplane at night in smooth air, you
have no sensation of motion, even though the plane may be moving at 800 km>h (500 mi>h). Why?
Q4.5 If the two ends of a rope in equilibrium are pulled with
forces of equal magnitude and opposite directions, why isn’t the
total tension in the rope zero?
Q4.6 You tie a brick to the end of a rope and whirl the brick
around you in a horizontal circle. Describe the path of the brick
after you suddenly let go of the rope.
Q4.7 When a car stops suddenly, the passengers tend to move
forward relative to their seats. Why? When a car makes a sharp
turn, the passengers tend to slide to one side of the car. Why?
Q4.8 Some people say that the “force of inertia” (or “force of
momentum”) throws the passengers forward when a car brakes
sharply. What is wrong with this explanation?
Q4.9 A passenger in a moving bus with no windows notices that
a ball that has been at rest in the aisle suddenly starts to move toward the rear of the bus. Think of two possible explanations, and
devise a way to decide which is correct.
Q4.10 Suppose you chose the fundamental physical quantities
to be force, length, and time instead of mass, length, and time.
What would be the units of mass in terms of those fundamental
quantities?
Q4.11 Why is the earth only approximately an inertial reference
frame?
Q4.12 Does Newton’s second law hold true for an observer in a
van as it speeds up, slows down, or rounds a corner? Explain.
S
Q4.13 Some students refer to the quantity ma as “the force of
acceleration.” Is it correct to refer to this quantity as a force? If so,
what exerts this force? If not, what is a better description of this
quantity?
Q4.14 The acceleration of a falling body is measured in an elevator that is traveling upward at a constant speed of 9.8 m>s. What
value is obtained?
Q4.15 You can play catch with a softball in a bus moving with
constant speed on a straight road, just as though the bus were at
rest. Is this still possible when the bus is making a turn at constant
speed on a level road? Why or why not?
Q4.16 Students sometimes say that the force of gravity on an object is 9.8 m>s2. What is wrong with this view?
Q4.17 Why can it hurt your foot more to kick a big rock than a
small pebble? Must the big rock hurt more? Explain.
Q4.18 “It’s not the fall that hurts you; it’s the sudden stop at the
bottom.” Translate this saying into the language of Newton’s laws
of motion.
Q4.19 A person can dive into water from a height of 10 m without injury, but a person who jumps off the roof of a 10-m-tall
building and lands on a concrete street is likely to be seriously
injured. Why is there a difference?
Q4.20 Why are cars designed to crumple in front and back for
safety? Why not for side collisions and rollovers?
Q4.21 When a string barely strong enough lifts a heavy weight,
it can lift the weight by a steady pull; but if you jerk the string, it
will break. Explain in terms of Newton’s laws of motion.
Q4.22 A large crate is suspended from the end of a vertical rope.
Is the tension in the rope greater when the crate is at rest or when
it is moving upward at constant speed? If the crate is traveling upward, is the tension in the rope greater when the crate is speeding
up or when it is slowing down? In each case, explain in terms of
Newton’s laws of motion.
Q4.23 Which feels a greater pull due to the earth’s gravity: a
10-kg stone or a 20-kg stone? If you drop the two stones, why
doesn’t the 20-kg stone fall with twice the acceleration of the
10-kg stone? Explain.
Q4.24 Why is it incorrect to say that 1.0 kg equals 2.2 lb?
Q4.25 A horse is hitched to a wagon. Since the wagon pulls back
on the horse just as hard as the horse pulls on the wagon, why doesn’t
the wagon remain in equilibrium, no matter how hard the horse pulls?
Q4.26 True or false? You exert a push P on an object and it
pushes back on you with a force F. If the object is moving at constant velocity, then F is equal to P, but if the object is being accelerated, then P must be greater than F.
Q4.27 A large truck and a small compact car have aShead-on collision. During the collision, the Struck exerts a force FT on C on the
car, and the car exerts a force FC on T on the truck. Which force
has the larger magnitude, or are they the same? Does your answer
depend on how fast each vehicle was moving before the collision?
Why or why not?
Q4.28 When a car comes to a stop on a level highway, what force
causes it to slow down? When the car increases its speed on the
same highway, what force causes it to speed up? Explain.
Q4.29 A small compact car is pushing a large van that has broken down, and they travel along the road with equal velocities and
accelerations. While the car is speeding up, is the force it exerts
on the van larger than, smaller than, or the same magnitude as the
force the van exerts on it? Which vehicle has the larger net force
on it, or are the net forces the same? Explain.
Q4.30 Consider a tug-of-war between two people who pull in opposite directions on the ends of a rope. By Newton’s third law, the
force that A exerts on B is just as great as the force that B exerts
on A. So what determines who wins? (Hint: Draw a free-body diagram showing all the forces that act on each person.)
Q4.31 Boxes A and B are in Figure Q4.31
contact on a horizontal, frictionless surface. You push 100 N
A
on box A with a horizontal
B
100-N force (Fig. Q4.31). Box A
weighs 150 N, and box B
weighs 50 N. Is the force that
box A exerts on box B equal to 100 N, greater than 100 N, or less
than 100 N? Explain.
Q4.32 A manual for student pilots contains this passage: “When
an airplane flies at a steady altitude, neither climbing nor descending, the upward lift force from the wings equals the plane’s
weight. When the plane is climbing at a steady rate, the upward
124
Chapter 4 Newton’s Laws of Motion
lift is greater than the weight; when the plane is descending at
a steady rate, the upward lift is less than the weight.” Are these
statements correct? Explain.
Q4.33 If your hands are wet and no towel is handy, you can remove
some of the excess water by shaking them. Why does this work?
Q4.34 If you squat down (such as when you examine the books
on a bottom shelf) and then suddenly get up, you may temporarily
feel light-headed. What do Newton’s laws of motion have to say
about why this happens?
Q4.35 When a car is hit from behind, the occupants may experience whiplash. Use Newton’s laws of motion to explain what
causes this result.
Q4.36 In a head-on auto collision, passengers who are not wearing
seat belts may be thrown through the windshield. Use Newton’s
laws of motion to explain why this happens.
Q4.37 In a head-on collision between a compact 1000-kg car and a
large 2500-kg car, which one experiences the greater force? Explain.
Which one experiences the greater acceleration? Explain why. Why
are passengers in the small car more likely to be injured than those
in the large car, even when the two car bodies are equally strong?
Q4.38 Suppose you are in a rocket with no windows, traveling
in deep space far from other objects. Without looking outside the
rocket or making any contact with the outside world, explain how
you could determine whether the rocket is (a) moving forward at
a constant 80% of the speed of light and (b) accelerating in the
forward direction.
exerciSeS
Section 4.1 Force and Interactions
.. Two dogs pull horizontally on ropes attached to a post; the
angle between the ropes is 60.0°. If Rover exerts a force of 270 N
and Fido exerts a force of 300 N, find the magnitude of the resultant force and the angle it makes with Rover’s rope.
4.2 . To extricate an SUV stuck in the mud, workmen use three
horizontal ropes, producing the force vectors shown in Fig. E4.2.
(a) Find the x- and y-components of each of the three pulls.
(b) Use the components to find the magnitude and direction of
the resultant of the three pulls.
4.1
Figure E4.2
y
788 N
32°
985 N
31°
x
53°
411 N
. bio Jaw Injury. Due to a
jaw injury, a patient must wear a
strap (Fig. E4.3) that produces a net
upward force of 5.00 N on his chin.
The tension is the same throughout
the strap. To what tension must the
strap be adjusted to provide the necessary upward force?
4.3
Figure E4.3
75.0°
4.4 . A man is dragging a Figure E4.4
trunk up the loading ramp of a
mover’s truck. The ramp has a
slope angle of 20.0°, and the man
S
S
pulls upward with a force F
F
whose direction makes an
30.0°
angle of 30.0° with the ramp
(Fig. E4.4).
(a) How large a
S
20.0°
force F is necessary for the
component Fx parallel to the
ramp to be 90.0 N? (b) How
large will the component
Fy perpendicular to the ramp be then?
S
S
S
4.5 . Forces F1 and F2 act at a point. The magnitude of F1 is
9.00 N, and its direction is 60.0°
above the x-axis in the second
S
quadrant. The magnitude of F2 is 6.00 N, and its direction is 53.1°
below the x-axis in the third quadrant. (a) What are the x- and
y-components of the resultant force? (b) What is the magnitude of
the resultant force?
Section 4.3 Newton’s Second Law
4.6 . An electron (mass = 9.11 * 10-31 kg) leaves one end of a
TV picture tube with zero initial speed and travels in a straight
line to the accelerating grid, which is 1.80 cm away. It reaches the
grid with a speed of 3.00 * 106 m>s. If the accelerating force is
constant, compute (a) the acceleration; (b) the time to reach the
grid; and (c) the net force, in newtons. Ignore the gravitational
force on the electron.
4.7 .. A 68.5-kg skater moving initially at 2.40 m>s on rough
horizontal ice comes to rest uniformly in 3.52 s due to friction
from the ice. What force does friction exert on the skater?
4.8 .. You walk into an elevator, step onto a scale, and push the
“up” button. You recall that your normal weight is 625 N. Draw a
free-body diagram. (a) When the elevator has an upward acceleration of magnitude 2.50 m>s2, what does the scale read? (b) If you
hold a 3.85-kg package by a light vertical string, what will be the
tension in this string when the elevator accelerates as in part (a)?
4.9 . A box rests on a frozen pond, which serves as a frictionless horizontal surface. If a fisherman applies a horizontal force
with magnitude 48.0 N to the box and produces an acceleration of
magnitude 2.20 m>s2, what is the mass of the box?
4.10 .. A dockworker applies a constant horizontal force of
80.0 N to a block of ice on a smooth horizontal floor. The frictional force is negligible. The block starts from rest and moves
11.0 m in 5.00 s. (a) What is the mass of the block of ice? (b) If the
worker stops pushing at the end of 5.00 s, how far does the block
move in the next 5.00 s?
4.11 . A hockey puck with mass 0.160 kg is at rest at the origin
(x = 0) on the horizontal, frictionless surface of the rink. At time
t = 0 a player applies a force of 0.250 N to the puck, parallel to
the x-axis; she continues to apply this force until t = 2.00 s.
(a) What are the position and speed of the puck at t = 2.00 s?
(b) If the same force is again applied at t = 5.00 s, what are the
position and speed of the puck at t = 7.00 s?
4.12 . A crate with mass 32.5 kg initially at rest on a warehouse
floor is acted on by a net horizontal force of 14.0 N. (a) What acceleration is produced? (b) How far does the crate travel in 10.0 s?
(c) What is its speed at the end of 10.0 s?
4.13 . A 4.50-kg experimental cart undergoes an acceleration in a
straight line (the x-axis). The graph in Fig. E4.13 shows this acceleration as a function of time. (a) Find the maximum net force on this
cart. When does this maximum force occur? (b) During what times
problems
is the net force on the cart a constant? (c) When is the net force equal
to zero?
Figure E4.13
ax (m>s2)
10.0
5.0
O
4.14
2.0
4.0
6.0
t (s)
.
A 2.75-kg cat moves Figure E4.14
in a straight line (the x-axis).
vx (m>s)
Figure E4.14 shows a graph of 12.0
the x-component of this cat’s 10.0
8.0
velocity as a function of time.
6.0
(a) Find the maximum net
4.0
2.0
force on this cat. When does
t (s)
this force occur? (b) When is
O
2.0 4.0 6.0 8.0 10.0
the net force on the cat equal
to zero? (c) What is the net force at time 8.5 s?
4.15 . A small 8.00-kg rocket burns fuel that exerts a timevarying upward force on the rocket (assume constant mass) as the
rocket moves upward from the launch pad. This force obeys the
equation F = A + Bt 2. Measurements show that at t = 0, the
force is 100.0 N, and at the end of the first 2.00 s, it is 150.0 N.
(a) Find the constants A and B, including their SI units. (b) Find the
net force on this rocket and its acceleration (i) the instant after the
fuel ignites and (ii) 3.00 s after the fuel ignites. (c) Suppose that
you were using this rocket in outer space, far from all gravity.
What would its acceleration be 3.00 s after fuel ignition?
Section 4.4 Mass and Weight
. An astronaut’s pack weighs 17.5 N when she is on the
earth but only 3.24 N when she is at the surface of a moon.
(a) What is the acceleration due to gravity on this moon?
(b) What is the mass of the pack on this moon?
4.17 . Superman throws a 2400-N boulder at an adversary.
What horizontal force must Superman apply to the boulder to give
it a horizontal acceleration of 12.0 m>s2?
4.18 . bio (a) An ordinary flea has a mass of 210 mg. How
many newtons does it weigh? (b) The mass of a typical froghopper
is 12.3 mg. How many newtons does it weigh? (c) A house cat
typically weighs 45 N. How many pounds does it weigh, and what
is its mass in kilograms?
4.19 . At the surface of Jupiter’s moon Io, the acceleration due to
gravity is g = 1.81 m>s2. A watermelon weighs 44.0 N at the
surface of the earth. (a) What is the watermelon’s mass on the earth’s
surface? (b) What would be its mass and weight on the surface of Io?
4.16
Section 4.5 Newton’s Third Law
4.20
.
A small car of mass 380 kg is pushing a large truck of
mass 900 kg due east on a level road. The car exerts a horizontal
force of 1600 N on the truck. What is the magnitude of the force
that the truck exerts on the car?
4.21 . bio World-class sprinters can accelerate out of the starting blocks with an acceleration that is nearly horizontal and has
magnitude 15 m>s2. How much horizontal force must a 55-kg
sprinter exert on the starting blocks to produce this acceleration?
Which body exerts the force that propels the sprinter: the blocks
or the sprinter herself?
125
4.22 .. The upward normal force exerted by the floor is 620 N
on an elevator passenger who weighs 650 N. What are the reaction
forces to these two forces? Is the passenger accelerating? If so,
what are the magnitude and direction of the acceleration?
4.23 .. Boxes A and B are in contact on a horizontal, frictionless
surface (Fig. E4.23). Box A has mass 20.0 kg and box B has mass
5.0 kg. A horizontal force of 250 N is exerted on box A. What is
the magnitude of the force that box A exerts on box B?
Figure E4.23
250 N
A
B
4.24
..
A student of mass 45 kg jumps off a high diving board.
What is the acceleration of the earth toward her as she accelerates toward the earth with an acceleration of 9.8 m>s2? Use
6.0 * 1024 kg for the mass of the earth, and assume that the net
force on the earth is the force of gravity she exerts on it.
Section 4.6 Free-Body Diagrams
4.25 .. Crates A and B sit at rest side by side on a frictionless
horizontal surface. They have
masses m A and m B, respectively.
S
When a horizontal force F is applied to crate A, the two crates
move off to the right. (a) Draw clearly labeled free-body diagrams
for crate A and for crate B. Indicate which pairs of forces, if any,
S
are third-law action–reaction pairs. (b) If the magnitude of F is
less than the total weight of the two crates, will it cause the crates
to move? Explain.
4.26 .. You pull horizontally Figure E4.26
on block B in Fig. E4.26, causA
ing both blocks to move together as a unit. For this
B
Pull
moving system, make a carefully labeled free-body diaHorizontal table
gram of block A if (a) the table
is frictionless and (b) there is friction between block B and the
table and the pull is equal in magnitude to the friction force on
block B due to the table.
4.27 . A ball is hanging from a long string that is tied to the ceiling of a train car traveling eastward on horizontal tracks. An observer inside the train car sees the ball hang motionless. Draw a
clearly labeled free-body diagram for the ball if (a) the train has a
uniform velocity and (b) the train is speeding up uniformly. Is the
net force on the ball zero in either case? Explain.
4.28 .. CP A .22-caliber rifle bullet traveling at 350 m>s strikes
a large tree and penetrates it to a depth of 0.130 m. The mass of
the bullet is 1.80 g. Assume a constant retarding force. (a) How
much time is required for the bullet to stop? (b) What force, in
newtons, does the tree exert on the bullet?
4.29 .. A chair of mass 12.0 kg is sitting on the horizontal floor;
the floor is not frictionless. You push on the chair with a force
F = 40.0 N that is directed at an angle of 37.0° below the horizontal, and the chair slides along the floor. (a) Draw a clearly labeled free-body diagram for the chair. (b) Use your diagram and
Newton’s laws to calculate the normal force that the floor exerts
on the chair.
126
Chapter 4 Newton’s Laws of Motion
probleMS
... A large box containing your new computer sits on the
bed of your pickup truck. You are stopped at a red light. When the
light turns green, you stomp on the gas and the truck accelerates.
To your horror, the box starts to slide toward the back of the truck.
Draw clearly labeled free-body diagrams for the truck and for the
box. Indicate pairs of forces, if any, that are third-law action–
reaction pairs. (The horizontal truck bed is not frictionless.)
4.31 .. CP A 5.60-kg bucket of water is accelerated upward by a
cord of negligible mass whose breaking strength is 75.0 N. If the
bucket starts from rest, what is the minimum time required to raise
the bucket a vertical distance of 12.0 m without breaking the cord?
4.32 .. CP You have just landed on Planet X. You release a
100-g ball from rest from a height of 10.0 m and measure that it
takes 3.40 s to reach the ground. Ignore any force on the ball from
the atmosphere of the planet. How much does the 100-g ball weigh
on the surface of Planet X?
4.33 .. Two adults and a child Figure P4.33
want to push a wheeled cart in
the direction marked x in
Fig. P4.33. The two adults
push
S
S
F1 = 100 N
with horizontal forces F1 and F2
as shown. (a) Find the magni60°
tude and direction of the smallx
est force that the child should
30°
exert. Ignore the effects of friction. (b) If the child exerts the
minimum force found in part (a),
F2 = 140 N
the cart accelerates at 2.0 m>s2 in
the + x-direction. What is the weight of the cart?
4.34 . CP An oil tanker’s engines have broken down, and the
wind is blowing the tanker straight toward a reef at a constant
speed of 1.5 m>s (Fig. P4.34). When the tanker is 500 m from the
reef, the wind dies down just as the engineer gets the engines
going again. The rudder is stuck, so the only choice is to try to accelerate straight backward away from the reef. The mass of the
tanker and cargo is 3.6 * 107 kg, and the engines produce a net
horizontal force of 8.0 * 104 N on the tanker. Will the ship hit
the reef? If it does, will the oil be safe? The hull can withstand an
impact at a speed of 0.2 m>s or less. Ignore the retarding force of
the water on the tanker’s hull.
4.30
Figure P4.34
F =
8.0 * 104 N
4.35
Figure P4.39
6.00 kg
4.00 kg
4.40
T
F
..
CP Two blocks connected by a light horizontal rope sit at
rest on a horizontal, frictionless surface. Block A has mass 15.0
kg, and block B has mass m. A constant horizontal force
F = 60.0 N is applied to block A (Fig. P4.40). In the first 5.00 s
after the force is applied, block A moves 18.0 m to the right.
(a) While the blocks are moving, what is the tension T in the rope
that connects the two blocks? (b) What is the mass of block B?
Figure P4.40
A
v = 1.5 m>s
3.6 * 107 kg
..
stop a 850-kg automobile traveling initially at 45.0 km>h in a distance equal to the diameter of a dime, 1.8 cm?
4.37 .. bio Human Biomechanics. The fastest pitched baseball
was measured at 46 m>s. A typical baseball has a mass of 145 g. If
the pitcher exerted his force (assumed to be horizontal and constant)
over a distance of 1.0 m, (a) what force did he produce on the ball
during this record-setting pitch? (b) Draw free-body diagrams of
the ball during the pitch and just after it left the pitcher’s hand.
4.38 .. bio Human Biomechanics. The fastest served tennis
ball, served by “Big Bill” Tilden in 1931, was measured at
73.14 m>s. The mass of a tennis ball is 57 g, and the ball, which
starts from rest, is typically in contact with the tennis racquet for
30.0 ms. Assuming constant acceleration, (a) what force did Big
Bill’s tennis racquet exert on the ball if he hit it essentially horizontally? (b) Draw free-body diagrams of the ball during the
serve and just after it moved free of the racquet.
4.39 . Two crates, one with mass 4.00 kg and the other with
mass 6.00 kg, sit on the frictionless surface of a frozen pond, connected by a light rope (Fig. P4.39). A woman wearing golf shoes
(for traction) pulls horizontally on the 6.00-kg crate with a force F
that gives the crate an acceleration of 2.50 m>s2. (a) What is the
acceleration of the 4.00-kg crate? (b) Draw a free-body diagram
for the 4.00-kg crate. Use that diagram and Newton’s second law
to find the tension T in the rope that connects the two crates.
(c) Draw a free-body diagram for the 6.00-kg crate. What is the
direction of the net force on the 6.00-kg crate? Which is larger in
magnitude, T or F? (d) Use part (c) and Newton’s second law to
calculate the magnitude of F.
F
B
500 m
CP bio A Standing Vertical Jump. Basketball player
Darrell Griffith is on record as attaining a standing vertical jump
of 1.2 m (4 ft). (This means that he moved upward by 1.2 m after
his feet left the floor.) Griffith weighed 890 N (200 lb). (a) What
was his speed as he left the floor? (b) If the time of the part of the
jump before his feet left the floor was 0.300 s, what was his average acceleration (magnitude and direction) while he pushed
against the floor? (c) Draw his free-body diagram. In terms of the
forces on the diagram, what was the net force on him? Use
Newton’s laws and the results of part (b) to calculate the average
force he applied to the ground.
4.36 ... CP An advertisement claims that a particular automobile can “stop on a dime.” What net force would be necessary to
4.41
.
CALC To study damage to aircraft that collide with large
birds, you design a test gun that will accelerate chicken-sized objects so that their displacement along the gun barrel is given by
x = 19.0 * 103 m>s22t 2 - 18.0 * 104 m>s32t 3. The object leaves
the end of the barrel at t = 0.025 s. (a) How long must the gun
barrel be? (b) What will be the speed of the objects as they leave
the end of the barrel? (c) What net force must be exerted on a
1.50-kg object at (i) t = 0 and (ii) t = 0.025 s?
4.42 .. CP A 6.50-kg instrument is hanging by a vertical wire
inside a spaceship that is blasting off from rest at the earth’s surface. This spaceship reaches an altitude of 276 m in 15.0 s with
constant acceleration. (a) Draw a free-body diagram for the instrument during this time. Indicate which force is greater. (b) Find the
force that the wire exerts on the instrument.
problems
4.43 .. bio Insect Dynamics. The froghopper (Philaenus
spumarius), the champion leaper of the insect world, has a mass of
12.3 mg and leaves the ground (in the most energetic jumps) at
4.0 m>s from a vertical start. The jump itself lasts a mere 1.0 ms
before the insect is clear of the ground. Assuming constant acceleration, (a) draw a free-body diagram of this mighty leaper during
the jump; (b) find the force that the ground exerts on the froghopper during the jump; and (c) express the force in part (b) in
terms of the froghopper’s weight.
4.44 . A loaded elevator with very worn cables has a total mass
of 2200 kg, and the cables can withstand a maximum tension of
28,000 N. (a) Draw the free-body force diagram for the elevator.
In terms of the forces on your diagram, what is the net force on the
elevator? Apply Newton’s second law to the elevator and find the
maximum upward acceleration for the elevator if the cables are
not to break. (b) What would be the answer to part (a) if the elevator were on the moon, where g = 1.62 m>s2?
4.45 .. CP After an annual checkup, you leave your physician’s
office, where you weighed 683 N. You then get into an elevator that,
conveniently, has a scale. Find the magnitude and direction of the
elevator’s acceleration if the scale reads (a) 725 N and (b) 595 N.
4.46 ... CP A nail in a pine board stops a 4.9-N hammer head from
an initial downward velocity of 3.2 m>s in a distance of 0.45 cm. In
addition, the person using the hammer exerts a 15-N downward force
on it. Assume that the acceleration of the hammer head is constant
while it is in contact with the nail and moving downward. (a) Draw
a free-body diagram for the hammer head. Identify the reaction
force for each
action force in the diagram. (b) Calculate the downS
ward force F exerted by the hammer head on the nail while the
hammer head is in contact with the nail and moving downward.
(c) Suppose that the nail is in hardwood and the distance the hammer head travels in coming to rest is only 0.12 cm. The downward
forces on theShammer head are the same as in part (b). What then
is the force F exerted by the hammer head on the nail while the
hammer head is in contact with the nail and moving downward?
4.47 .. CP Jumping to the Ground. A 75.0-kg man steps off
a platform 3.10 m above the ground. He keeps his legs straight as
he falls, but his knees begin to bend at the moment his feet touch
the ground; treated as a particle, he moves an additional 0.60 m
before coming to rest. (a) What is his speed at the instant his feet
touch the ground? (b) If we treat the man as a particle, what is his
acceleration (magnitude and direction) as he slows down, if
the acceleration is assumed to be constant? (c) Draw his freebody diagram. In terms of the forces on the diagram, what is the
net force on him? Use Newton’s laws and the results of part (b)
to calculate the average force his feet exert
on the ground while he slows down.
Figure P4.48
Express this force both in newtons and
as a multiple of his weight.
F = 200 N
4.48 .. The two blocks in Fig. P4.48
are connected by a heavy uniform rope
with a mass of 4.00 kg. An upward force
6.00 kg
of 200 N is applied as shown. (a) Draw
three free-body diagrams: one for the
6.00-kg block, one for the 4.00-kg rope,
and another one for the 5.00-kg block.
4.00 kg
For each force, indicate what body exerts that force. (b) What is the acceleration of the system? (c) What is the
tension at the top of the heavy rope?
5.00 kg
(d) What is the tension at the midpoint
of the rope?
127
4.49 .. CP Boxes A and B are connected to each end of a light
vertical rope (Fig. P4.49). A constant upward force F = 80.0 N is
applied to box A. Starting from rest, box B descends 12.0 m in
4.00 s. The tension in the rope connecting the two boxes is 36.0 N.
What are the masses of (a) box B, (b) box A?
Figure P4.49
F
A
B
4.50
..
CP Extraterrestrial Physics. You have landed on an
unknown planet, Newtonia, and want to know what objects weigh
there. When you push a certain tool, starting from rest, on a frictionless horizontal surface with a 12.0-N force, the tool moves
16.0 m in the first 2.00 s. You next observe that if you release this
tool from rest at 10.0 m above the ground, it takes 2.58 s to reach
the ground. What does the tool weigh on Newtonia, and what does
it weigh on Earth?
4.51 .. CP CALC A mysterious rocket-propelled object of mass
45.0 kg is initially at rest in the middle of the horizontal, frictionless surface of an ice-covered lake. Then a force directed east and
with magnitude F(t) = (16.8 N>s)t is applied. How far does the
object travel in the first 5.00 s after the force is applied?
4.52 ... CALC The position of a training helicopter (weight
2.75 * 105 N) in a test is given by nr = (0.020 m>s3)t 3nd +
(2.2 m>s) ten − (0.060 m>s2)t 2kn . Find the net force on the helicopter
at t = 5.0 s.
4.53 .. DATA The table* gives automobile performance data for
a few types of cars:
Make and Model (Year)
Mass (kg)
Time (s) to go from
0 to 60 mph
Alpha Romeo 4C (2013)
895
4.4
Honda Civic 2.0i (2011)
1320
6.4
Ferrari F430 (2004)
1435
3.9
Ford Focus RS500 (2010)
1468
5.4
Volvo S60 (2013)
1650
7.2
*Source: www.autosnout.com
(a) During an acceleration of 0 to 60 mph, which car has the
largest average net force acting on it? The smallest? (b) During
this acceleration, for which car would the average net force on a
72.0-kg passenger be the largest? The smallest? (c) When the
Ferrari F430 accelerates from 0 to 100 mph in 8.6 s, what is the
average net force acting on it? How does this net force compare
with the average net force during the acceleration from 0 to
60 mph? Explain why these average net forces might differ.
(d) Discuss why a car has a top speed. What is the net force on the
Ferrari F430 when it is traveling at its top speed, 196 mph?
4.54 .. DATA An 8.00-kg box sits on a level floor. You give the
box a sharp push and find that it travels 8.22 m in 2.8 s before
coming to rest again. (a) You measure that with a different push
the box traveled 4.20 m in 2.0 s. Do you think the box has a constant acceleration as it slows down? Explain your reasoning.
(b) You add books to the box to increase its mass. Repeating the
experiment, you give the box a push and measure how long it
takes the box to come to rest and how far the box travels. The
128
Chapter 4 Newton’s Laws of Motion
results, including the initial experiment with no added mass, are
given in the table:
Distance (m)
Time (s)
8.22
2.8
3.00
10.75
3.2
7.00
9.45
3.0
7.10
2.6
0
12.0
In each case, did your push give the box the same initial speed?
What is the ratio between the greatest initial speed and the smallest initial speed for these four cases? (c) Is the average horizontal
force f exerted on the box by the floor the same in each case?
Graph the magnitude of force f versus the total mass m of the box
plus its contents, and use your graph to determine an equation for f
as a function of m.
4.55 .. DATA You are a Starfleet captain going boldly where no
man has gone before. You land on a distant planet and visit an engineering testing lab. In one experiment a short, light rope is attached to the top of a block and a constant upward force F is
applied to the free end of the rope. The block has mass m and is
initially at rest. As F is varied, the time for the block to move upward 8.00 m is measured. The values that you collected are given
in the table:
F (N)
Time (s)
250
300
350
400
450
500
3.3
2.2
1.7
1.5
1.3
1.2
Figure P4.60
1600
(a) Plot F versus the acceleration a of the block. (b) Use your graph
to determine the mass m of the block and the acceleration of gravity g at the surface of the planet. Note that even on that planet,
measured values contain some experimental error.
challenGe problem
4.56
1400
1200
1000
800
600
400
...
CALC An object ofSmass m is at rest in equilibrium at the
origin. At t = 0 a new force F(t) is applied that has components
Fx(t) = k1 + k2y
bio Forces on a Dancer’s BoDy. Dancers experience
large forces associated with the jumps they make. For example,
when a dancer lands after a vertical jump, the force exerted on the
head by the neck must exceed the head’s weight by enough to
cause the head to slow down and come to rest. The head is about
9.4% of a typical person’s mass. Video analysis of a 65-kg dancer
landing after a vertical jump shows that her head decelerates from
4.0 m>s to rest in a time of 0.20 s.
4.57 What is the magnitude of the average force that her neck
exerts on her head during the landing? (a) 0 N; (b) 60 N; (c) 120 N;
(d) 180 N.
4.58 Compared with the force her neck exerts on her head during
the landing, the force her head exerts on her neck is (a) the same;
(b) greater; (c) smaller; (d) greater during the first half of the landing and smaller during the second half of the landing.
4.59 While the dancer is in the air and holding a fixed pose,
what is the magnitude of the force her neck exerts on her head?
(a) 0 N; (b) 60 N; (c) 120 N; (d) 180 N.
4.60 The forces on a dancer can be measured directly when a
dancer performs a jump on a force plate that measures the force
between her feet and the ground. A graph of force versus time
throughout a vertical jump performed on a force plate is shown in
Fig. P4.60. What is happening at 0.4 s? The dancer is (a) bending
her legs so that her body is accelerating downward; (b) pushing
her body up with her legs and is almost ready to leave the ground;
(c) in the air and at the top of her jump; (d) landing and her feet
have just touched the ground.
Force (N)
Added Mass (kg)
paSSaGe problemS
Fy(t) = k3t
S
where k1, k2, and k3 are constants. Calculate the position r (t) and
S
velocity v(t) vectors as functions of time.
200
0
- 200
0.5
1.0
1.5
2.0
Time (s)
2.5
3.0
answers
129
Answers
chapter opening Question
?
(v) Newton’s third law tells us that the barbell pushes on the
weightlifter just as hard as the weightlifter pushes on the barbell in all circumstances, no matter how the barbell is moving.
However, the magnitude of the force that the weightlifter exerts
is different in different circumstances. This force magnitude is
equal to the weight of the barbell when the barbell is stationary,
moving upward at a constant speed, or moving downward at a
constant speed; it is greater than the weight of the barbell when
the barbell accelerates upward; and it is less than the weight of
the barbell when the barbell accelerates downward. But in each
case the push of the barbell on the weightlifter has exactly the
same magnitude as the push of the weightlifter on the barbell.
test your understanding Questions
4.1 (iv) The gravitational force on the crate points straight
downward. In Fig. 4.5 the x-axis points up and to the right,
and the y-axis points up and to the left. Hence the gravitational
force has both an x-component and a y-component, and both are
negative.
4.2 (i), (ii), and (iv) In (i), (ii), and (iv) the body is not accelerating, so the net force on the body is zero. [In (iv), the box
remains stationary as seen in the inertial reference frame of the
ground as the truck accelerates forward, like the person on skates
in Fig. 4.10a.] In (iii), the hawk is moving in a circle; hence it is
accelerating and is not in equilibrium.
4.3 (iii), (i) and (iv) (tie), (ii) The acceleration is equal to the
net force divided by the mass. Hence the magnitude of the acceleration in each situation is
(i) a = 12.0 N2>12.0 kg2 = 1.0 m>s2;
(ii) a = 18.0 N2>12.0 N2 = 4.0 m>s2;
(iii) a = 12.0 N2>18.0 kg2 = 0.25 m>s2;
(iv) a = 18.0 N2>18.0 kg2 = 1.0 m>s2.
4.4 It would take twice the effort for the astronaut to walk
around because her weight on the planet would be twice as much
as on the earth. But it would be just as easy to catch a ball moving horizontally. The ball’s mass is the same as on earth, so the
horizontal force the astronaut would have to exert to bring it to a
stop (i.e., to give it the same acceleration) would also be the same
as on earth.
4.5 By Newton’s third law, the two forces have equal magnitude.
Because the car has much greater mass than the mosquito, it undergoes only a tiny, imperceptible acceleration in response to the
force of the impact. By contrast, the mosquito, with its minuscule
mass, undergoes a catastrophically large acceleration.
4.6 (iv) The buoyancy force is an upward force that the water
exerts on the swimmer. By Newton’s third law, the other half of
the action–reaction pair is a downward force that the swimmer
exerts on the water and has the same magnitude as the buoyancy
force. It’s true that the weight of the swimmer is also downward
and has the same magnitude as the buoyancy force; however, the
weight acts on the same body (the swimmer) as the buoyancy
force, and so these forces aren’t an action–reaction pair.
bridging problem
(a) See the Video Tutor Solution on MasteringPhysics®
(b) (i) 2.20 m>s2; (ii) 6.00 N; (iii) 3.00 N
?
Each of the seeds being
blown off the head of a
dandelion (genus Taraxacum)
has a feathery structure called a
pappus. The pappus acts like a
parachute and enables the seed
to be borne by the wind and
drift gently to the ground. If a
seed with its pappus descends
straight down at a steady
speed, which force acting on
the seed has a greater magnitude? (i) The force of gravity;
(ii) the upward force exerted
by the air; (iii) both forces have
the same magnitude; (iv) it
depends on the speed at
which the seed descends.
5
Applying
newton’s lAws
Learning goaLs
Looking forward at …
5.1 How to use Newton’s first law to solve prob-
5.2
5.3
5.4
5.5
lems involving the forces that act on a body
in equilibrium.
How to use Newton’s second law to solve
problems involving the forces that act on an
accelerating body.
The nature of the different types of friction
forces—static friction, kinetic friction, rolling
friction, and fluid resistance—and how to
solve problems that involve these forces.
How to solve problems involving the forces
that act on a body moving along a circular
path.
The key properties of the four fundamental
forces of nature.
Looking back at …
1.8 Determining the components of a vector
from its magnitude and direction.
2.4 Straight-line motion with constant
3.3
3.4
4.1
4.2
4.3
acceleration.
Projectile motion.
Uniform and nonuniform circular motion.
Superposition of forces.
Newton’s first law.
Newton’s second law.
4.4 Mass and weight.
4.5 Newton’s third law.
.
130
W
e saw in Chapter 4 that Newton’s three laws of motion, the foundation
of classical mechanics, can be stated very simply. But applying these
laws to situations such as an iceboat skating across a frozen lake, a
toboggan sliding down a hill, or an airplane making a steep turn requires analytical skills and problem-solving technique. In this chapter we’ll help you extend
the problem-solving skills you began to develop in Chapter 4.
We’ll begin with equilibrium problems, in which we analyze the forces that
act on a body that is at rest or moving with constant velocity. We’ll then consider
bodies that are not in equilibrium, for which we’ll have to deal with the relationship between forces and motion. We’ll learn how to describe and analyze the
contact force that acts on a body when it rests on or slides over a surface. We’ll
also analyze the forces that act on a body that moves in a circle with constant
speed. We close the chapter with a brief look at the fundamental nature of force
and the classes of forces found in our physical universe.
5.1 Using newton’s First Law:
ParticLes in eqUiLibriUm
We learned in Chapter 4 that a body is in equilibrium when it is at rest or moving with constant velocity in an inertial frame of reference. A hanging lamp, a
kitchen table, an airplane flying straight and level at a constant speed—all are
examples of equilibrium situations. In this section we consider only the equilibrium of a body that can be modeled as a particle. (In Chapter 11 we’ll see how to
analyze a body in equilibrium that can’t be represented adequately as a particle,
such as a bridge that’s supported at various points along its span.) The essential
physical principle is Newton’s first law:
5.1 Using Newton’s First Law: Particles in Equilibrium
Newton’s first law:
Net force on a body ...
gF = 0
S
Sum of x-components of force
on body must be zero.
gFx = 0
131
... must be zero for a
body in equilibrium.
Sum of y-components of force
on body must be zero.
g Fy = 0
(5.1)
This section is about using Newton’s first law to solve problems dealing with
bodies in equilibrium. Some of these problems may seem complicated, but remember that all problems involving particles in equilibrium are done in the same
way. Problem-Solving Strategy 5.1 details the steps you need to follow for any
and all such problems. Study this strategy carefully, look at how it’s applied in
the worked-out examples, and try to apply it when you solve assigned problems.
Problem-Solving Strategy 5.1
newton’S FirSt law: equilibrium oF a Particle
identiFy the relevant concepts: You must use Newton’s first law,
Eqs. (5.1), for any problem that involves forces acting on a body in
equilibrium—that is, either at rest or moving with constant velocity.
A car is in equilibrium when it’s parked, but also when it’s traveling
down a straight road at a steady speed.
If the problem involves more than one body and the bodies interact with each other, you’ll also need to use Newton’s third law.
This law allows you to relate the force that one body exerts on a
second body to the force that the second body exerts on the first
one.
Identify the target variable(s). Common target variables in
equilibrium problems include the magnitude and direction (angle)
of one of the forces, or the components of a force.
set UP the problem by using the following steps:
1. Draw a very simple sketch of the physical situation, showing
dimensions and angles. You don’t have to be an artist!
2. Draw a free-body diagram for each body that is in equilibrium.
For now, we consider the body as a particle, so you can represent it as a large dot. In your free-body diagram, do not include
the other bodies that interact with it, such as a surface it may
be resting on or a rope pulling on it.
3. Ask yourself what is interacting with the body by contact or in
any other way. On your free-body diagram, draw a force vector
for each interaction. Label each force with a symbol for the
magnitude of the force. If you know the angle at which a force
is directed, draw the angle accurately and label it. Include the
body’s weight, unless the body has negligible mass. If the mass
is given, use w = mg to find the weight. A surface in contact
with the body exerts a normal force perpendicular to the surface and possibly a friction force parallel to the surface. A rope
or chain exerts a pull (never a push) in a direction along its
length.
4. Do not show in the free-body diagram any forces exerted by the
body on any other body. The sums in Eqs. (5.1) include only
forces that act on the body. For each force on the body, ask yourself “What other body causes that force?” If you can’t answer
that question, you may be imagining a force that isn’t there.
5. Choose a set of coordinate axes and include them in your freebody diagram. (If there is more than one body in the problem,
choose axes for each body separately.) Label the positive direction for each axis. If a body rests or slides on a plane surface,
for simplicity choose axes that are parallel and perpendicular
to this surface, even when the plane is tilted.
execUte the solution as follows:
1. Find the components of each force along each of the body’s
coordinate axes. Draw a wiggly line through each force vector
that has been replaced by its components, so you don’t count it
twice. The magnitude of a force is always positive, but its components may be positive or negative.
2. Set the sum of all x-components of force equal to zero. In a
separate equation, set the sum of all y-components equal to
zero. (Never add x- and y-components in a single equation.)
3. If there are two or more bodies, repeat all of the above steps for
each body. If the bodies interact with each other, use Newton’s
third law to relate the forces they exert on each other.
4. Make sure that you have as many independent equations as the
number of unknown quantities. Then solve these equations to
obtain the target variables.
evaLUate your answer: Look at your results and ask whether
they make sense. When the result is a symbolic expression or formula, check to see that your formula works for any special cases
(particular values or extreme cases for the various quantities) for
which you can guess what the results ought to be.
A gymnast with mass m G = 50.0 kg suspends herself from the
lower end of a hanging rope of negligible mass. The upper end
of the rope is attached to the gymnasium ceiling. (a) What is the
gymnast’s weight? (b) What force (magnitude and direction) does
the rope exert on her? (c) What is the tension at the top of the
rope?
Solution
examPle 5.1 one-dimenSional equilibrium: tenSion in a maSSleSS roPe
soLUtion
identiFy and set UP: The gymnast and the rope are in equilibrium, so we can apply Newton’s first law to both bodies. We’ll
use Newton’s third law to relate the forces that they exert on each
Continued
132
ChaPtEr 5 applying Newton’s Laws
other. The target variables are the gymnast’s weight, wG; the force
that the bottom of the rope exerts on the gymnast 1call it TR on G2;
and the force that the ceiling exerts on the top of the rope
1call it TC on R2. Figure 5.1 shows our sketch of the situation and
free-body diagrams for the gymnast and for the rope. We take the
positive y-axis to be upward in each diagram. Each force acts in
the vertical direction and so has only a y-component.
The forces TR on G (the upward force of the rope on the gymnast, Fig. 5.1b) and TG on R (the downward force of the gymnast
on the rope, Fig. 5.1c) form an action–reaction pair. By Newton’s
third law, they must have the same magnitude.
Note that Fig. 5.1c includes only the forces that act on the rope.
In particular, it doesn’t include the force that the rope exerts on
the ceiling (compare the discussion of the apple in Conceptual
Example 4.9 in Section 4.5).
5.1 Our sketches for this problem.
(a) The situation
(b) Free-body
diagram for gymnast
(c) Free-body
diagram for rope
execUte: (a) The magnitude of the gymnast’s weight is the prod-
uct of her mass and the acceleration due to gravity, g:
wG = m G g = 150.0 kg219.80 m>s22 = 490 N
(b) The gravitational force on the gymnast (her weight) points
in the negative y-direction, so its y-component is - wG. The upward force of the rope on the gymnast has unknown magnitude
TR on G and positive y-component +TR on G. We find this by using
Newton’s first law from Eqs. (5.1):
gFy = TR on G + 1- wG2 = 0
Gymnast:
so
TR on G = wG = 490 N
The rope pulls up on the gymnast with a force TR on G of magnitude 490 N. (By Newton’s third law, the gymnast pulls down on
the rope with a force of the same magnitude, TG on R = 490 N.)
(c) We have assumed that the rope is weightless, so the only
forces on it are those exerted by the ceiling (upward force of unknown magnitude TC on R) and by the gymnast (downward force
of magnitude TG on R = 490 N). From Newton’s first law, the net
vertical force on the rope in equilibrium must be zero:
Rope:
gFy = TC on R + 1- TG on R2 = 0
so
TC on R = TG on R = 490 N
evaLUate: The tension at any point in the rope is the magnitude
Action–
reaction
pair
Solution
examPle 5.2
of the force that acts at that point. For this weightless rope, the
tension TG on R at the lower end has the same value as the tension
TC on R at the upper end. For such an ideal weightless rope, the tension has the same value at any point along the rope’s length. (See
the discussion in Conceptual Example 4.10 in Section 4.5.)
one-dimenSional equilibrium: tenSion in a roPe with maSS
Find the tension at each end of the rope in Example 5.1 if the weight
of the rope is 120 N.
5.2 Our sketches for this problem, including the weight of the
rope.
(a) Free-body
diagram for gymnast
soLUtion
identiFy and set UP: As in Example 5.1, the target variables are
the magnitudes TG on R and TC on R of the forces that act at the bottom
and top of the rope, respectively. Once again, we’ll apply Newton’s
first law to the gymnast and to the rope, and use Newton’s third law to
relate the forces that the gymnast and rope exert on each other. Again
we draw separate free-body diagrams for the gymnast (Fig. 5.2a)
and the rope (Fig. 5.2b). There is now a third force acting on the rope,
however: the weight of the rope, of magnitude wR = 120 N.
(b) Free-body
diagram for rope
(c) Free-body diagram
for gymnast and rope
as a composite body
Action–
reaction
pair
execUte: The gymnast’s free-body diagram is the same as in
Example 5.1, so her equilibrium condition is also the same. From
Newton’s third law, TR on G = TG on R, and we again have
Gymnast:
gFy = TR on G + 1-wG2 = 0
so
TR on G = TG on R = wG = 490 N
The equilibrium condition gFy = 0 for the rope is now
Rope:
gFy = TC on R + 1- TG on R2 + 1-wR2 = 0
Note that the y-component of TC on R is positive because it points
in the + y@direction, but the y-components of both TG on R and
wR are negative. We solve for TC on R and substitute the values
TG on R = TR on G = 490 N and wR = 120 N:
TC on R = TG on R + wR = 490 N + 120 N = 610 N
evaLUate: When we include the weight of the rope, the tension
is different at the rope’s two ends: 610 N at the top and 490 N at
the bottom. The force TC on R = 610 N exerted by the ceiling has
to hold up both the 490-N weight of the gymnast and the 120-N
weight of the rope.
To see this more clearly, we draw a free-body diagram for a
composite body consisting of the gymnast and rope together
(Fig. 5.2c). Only two external forces act on this composite body:
the force TC on R exerted by the ceiling and the total weight
5.1 Using Newton’s First Law: Particles in Equilibrium
wG + wR = 490 N + 120 N = 610 N. (The forces TG on R and
TR on G are internal to the composite body. Newton’s first law applies only to external forces, so these internal forces play no role.)
Hence Newton’s first law applied to this composite body is
Composite body:
gFy = TC on R + 3-1wG + wR24 = 0
133
Treating the gymnast and rope as a composite body is simpler,
but we can’t find the tension TG on R at the bottom of the rope by
this method. Moral: Whenever you have more than one body in a
problem involving Newton’s laws, the safest approach is to treat
each body separately.
and so TC on R = wG + wR = 610 N.
two-dimenSional equilibrium
In Fig. 5.3a, a car engine with weight w hangs from a chain that
is linked at ring O to two other chains, one fastened to the ceiling and the other to the wall. Find expressions for the tension in
each of the three chains in terms of w. The weights of the ring and
chains are negligible compared with the weight of the engine.
soLUtion
identiFy and set UP: The target variables are the tension magnitudes T1, T2, and T3 in the three chains (Fig. 5.3a). All the bodies
are in equilibrium, so we’ll use Newton’s first law. We need three
independent equations, one for each target variable. However,
applying Newton’s first law in component form to just one body
gives only two equations [the x- and y-equations in Eqs. (5.1)]. So
we’ll have to consider more than one body in equilibrium. We’ll
look at the engine (which is acted on by T1) and the ring (which is
attached to all three chains and so is acted on by all three tensions).
Figures 5.3b and 5.3c show our free-body diagrams and choice
of coordinate axes. Two forces act on the engine: its weight w and
the upward force T1 exerted by the vertical chain. Three forces act
on the ring: the tensions from the vertical chain 1T12, the horizontal chain 1T22, and the slanted chain 1T32. Because the vertical
chain has negligible weight, it exerts forces of the same magnitude T1 at both of its ends (see Example 5.1). (If the weight of this
chain were not negligible, these two forces would have different
magnitudes; see Example 5.2.) The weight of the ring is also negligible, so it isn’t included in Fig. 5.3c.
execUte: The forces acting on the engine are along the y-axis
only, so Newton’s first law [Eqs. (5.1)] says
Engine:
Solution
examPle 5.3
gFy = T1 + 1- w2 = 0
and
T1 = w
The horizontal and slanted chains don’t exert forces on the
engine itself because they are not attached to it. These forces do
appear when we apply Newton’s first law to the ring, however.
In the free-body diagram for the ring (Fig. 5.3c), remember that
T1, T2, and T3 are the magnitudes of the forces. We resolve the
force with magnitude T3 into its x- and y-components. Applying
Newton’s first law in component form to the ring gives us the two
equations
gFx = T3 cos 60° + 1- T22 = 0
gFy = T3 sin 60° + 1- T12 = 0
Ring :
Ring :
Because T1 = w (from the engine equation), we can rewrite the
second ring equation as
T3 =
T1
w
=
= 1.2w
sin 60°
sin 60°
We can now use this result in the first ring equation:
T2 = T3 cos 60° = w
cos 60°
= 0.58w
sin 60°
evaLUate: The chain attached to the ceiling exerts a force on the
ring with a vertical component equal to T1, which in turn is equal
to w. But this force also has a horizontal component, so its magnitude T3 is somewhat greater than w. This chain is under the greatest tension and is the one most susceptible to breaking.
To get enough equations to solve this problem, we had to consider not only the forces on the engine but also the forces acting
on a second body (the ring connecting the chains). Situations like
this are fairly common in equilibrium problems, so keep this technique in mind.
5.3 Our sketches for this problem.
(a) Engine, chains, and ring
60°
T2
O
T1
T3
(b) Free-body
diagram for engine
(c) Free-body
diagram for ring O
134
Chapter 5 applying Newton’s Laws
Solution
ExamplE 5.4
an inclinEd planE
A car of weight w rests on a slanted ramp attached to a trailer
(Fig. 5.4a). Only a cable running from the trailer to the car prevents the car from rolling off the ramp. (The car’s brakes are off
and its transmission is in neutral.) Find the tension in the cable and
the force that the ramp exerts on the car’s tires.
Solution
identify: The car is in equilibrium, so we use Newton’s first law.
The ramp exerts a separate force on each of the car’s tires, but for
simplicity we lump these forces into a single force. For a further
simplification, we’ll neglect any friction force the ramp exerts on
the tires (see Fig. 4.2b). Hence the ramp exerts only a force on the
car that is perpendicular to the ramp. As in Section 4.1, we call
this force the normal force (see Fig. 4.2a). The two target variables are the magnitude T of the tension in the cable and the magnitude n of the normal force.
Set up: Figure 5.4 shows the situation and a free-body diagram
for the car. The three forces acting on the car are its weight (magnitude w), the tension in the cable (magnitude T), and the normal
force (magnitude n). Note that the angle a between the ramp and
S
the horizontal is equal to the angle a between the weight vector w
and the downward normal to the plane of the ramp. Note also that
we choose the x- and y-axes to be parallel and perpendicular to the
ramp so that we need to resolve only one force (the weight) into
x- and y-components. If we had chosen axes that were horizontal
and vertical, we’d have to resolve both the normal force and the
tension into components.
execute: To write down the x- and y-components of Newton’s
first law, we must first find the components of the weight. One
complication is that the angle a in Fig. 5.4b is not measured from
the + x@axis toward the + y@axis. Hence we cannot use Eqs. (1.5)
directly to find the components. (You may want to review Section 1.8
to make sure that you understand this important point.)
5.4 A cable holds a car at rest on a ramp.
(a) Car on ramp
(b) Free-body diagram for car
We replace the weight
by its components.
y
n
n
T
x
w sin a
T
w cos a
a
w
a
w
gFx = T + 1-w sin a2 = 0
gFy = n + 1- w cos a2 = 0
(Remember that T, w, and n are all magnitudes of vectors and are
therefore all positive.) Solving these equations for T and n, we find
T = w sin a
n = w cos a
evaluate: Our answers for T and n depend on the value of a.
To check this dependence, let’s look at some special cases. If the
ramp is horizontal 1a = 02, we get T = 0 and n = w: No cable
tension T is needed to hold the car, and the normal force n is equal
in magnitude to the weight. If the ramp is vertical 1a = 90°2, we
get T = w and n = 0: The cable tension T supports all of the car’s
weight, and there’s nothing pushing the car against the ramp.
caution normal force and weight may not be equal It’s a common error to assume that the normal-force magnitude n equals the
weight w. Our result shows that this is not always the case. Always
treat n as a variable and solve for its value, as we’ve done here. ❙
?
How would the answers for T and n be affected if the car
were being pulled up the ramp at a constant speed? This, too,
is an equilibrium situation, since the car’s velocity is constant.
So the calculation is the same, and T and n have the same values
as when the car is at rest. (It’s true that T must be greater than
w sin a to start the car moving up the ramp, but that’s not what
we asked.)
Equilibrium of bodiEs connEctEd by cablE and pullEy
Your firm needs to haul granite blocks up a 15° slope out of a
quarry and to lower dirt into the quarry to fill the holes. You
design a system in which a granite block on a cart with steel
wheels (weight w1, including both block and cart) is pulled uphill
on steel rails by a dirt-filled bucket (weight w2, including both dirt
Solution
ExamplE 5.5
S
One way to find the components of w is to consider the right
triangles in Fig. 5.4b. The sine of a is the magnitude of the
S
x-component of w (that is, the side of the triangle opposite a)
divided by the magnitude w (the hypotenuse of the triangle).
Similarly, the cosine of a is the magnitude of the y-component
(the side of the triangle adjacent to a) divided by w. Both components are negative, so wx = - w sin a and wy = -w cos a.
S
Another approach is to recognize that one component of w
must involve sin a while the other component involves cos a. To
decide which is which, draw the free-body diagram so that the
angle a is noticeably smaller or larger than 45°. (You’ll have to
fight the natural tendency to draw such angles as being close to
45°.) We’ve drawn Fig. 5.4b so that a is smaller than 45°, so sin a
S
is less than cos a. The figure shows that the x-component of w is
smaller than the y-component, so the x-component must involve
sin a and the y-component must involve cos a. We again find
wx = - w sin a and wy = -w cos a.
In Fig. 5.4b we draw a wiggly line through the original vector
representing the weight to remind us not to count it twice. Newton’s
first law gives us
and bucket) that descends vertically into the quarry (Fig. 5.5a).
How must the weights w1 and w2 be related in order for the system to move with constant speed? Ignore friction in the pulley and
wheels, and ignore the weight of the cable.
5.2 Using Newton’s Second Law: Dynamics of Particles
135
5.5 Our sketches for this problem.
(a) Dirt-filled bucket pulls cart with granite block.
(d) Free-body
diagram for cart
Cart
(c) Free-body
diagram for bucket
Bucket
15°
(b) Idealized model of the system
execUte: Applying gFy = 0 to the bucket in Fig. 5.5c, we find
soLUtion
identiFy and set UP: The cart and bucket each move with a constant velocity (in a straight line at constant speed). Hence each
body is in equilibrium, and we can apply Newton’s first law to
each. Our target is an expression relating the weights w1 and w2.
Figure 5.5b shows our idealized model for the system, and
Figs. 5.5c and 5.5d show our free-body diagrams. The two forces
on the bucket are its weight w2 and an upward tension exerted by
the cable. As for the car on the ramp in Example 5.4, three forces
act on the cart: its weight w1, a normal force of magnitude n exerted by the rails, and a tension force from the cable. Since we’re
assuming that the cable has negligible weight, the tension forces
that the cable exerts on the cart and on the bucket have the same
magnitude T. (We’re ignoring friction, so we assume that the rails
exert no force on the cart parallel to the incline.) Note that we
orient the axes differently for each body; the choices shown are
the most convenient ones. We find the components of the weight
force in the same way that we did in Example 5.4. (Compare
Fig. 5.5d with Fig. 5.4b.)
gFy = T + 1-w22 = 0
gFx = T + 1-w1 sin 15°2 = 0
Each component of
net force on body ...
S
evaLUate: Our analysis doesn’t depend at all on the direction
in which the cart and bucket move. Hence the system can move
with constant speed in either direction if the weight of the dirt and
bucket is 26% of the weight of the granite block and cart. What
would happen if w2 were greater than 0.26w1? If it were less than
0.26w1?
Notice that we didn’t need the equation gFy = 0 for the cart
and block. Can you use this to show that n = w1 cos 15°?
... body has acceleration in
same direction as net force.
Mass of body
gFx = max
g Fy = may
T = w1 sin 15°
w2 = w1 sin 15° = 0.26w1
We are now ready to discuss dynamics problems. In these problems, we apply
Newton’s second law to bodies on which the net force is not zero. These bodies
are not in equilibrium and hence are accelerating:
S
so
Equating the two expressions for T, we find
5.2 Using newton’s second Law:
dynamics oF ParticLes
g F = ma
T = w2
Applying gFx = 0 to the cart (and block) in Fig. 5.5d, we get
test yoUr Understanding oF section 5.1 A traffic light of weight w
hangs from two lightweight cables, one on each side of the light. Each cable hangs at a
45° angle from the horizontal. What is the tension in each cable? (i) w>2; (ii) w>12;
(iii) w; (iv) w 12 ; (v) 2w. ❙
Newton’s second law:
If net force on a body
is not zero ...
so
(5.2)
... equals body’s mass
times corresponding
acceleration component.
(a)
w
136
ChaPtEr 5 applying Newton’s Laws
ay
Only the force of gravity
acts on this falling fruit.
RIGHT!
You can safely draw
the acceleration
vector to one side
of the diagram.
5.6 Correct and incorrect free-body diagrams for a falling body.
(b) Correct free-body diagram
(a)
(c) Incorrect free-body diagram
y
Only the force of gravity
acts on this falling fruit.
x
w
(b) Correct free-body diagram
y
x
w
ay
x
ma
w
Problem-Solving
ay
x
RIGHT!
You can safely draw
the acceleration
vector to one side
of the diagram.
ma
w
WRONG
This vector doesn’t
belong in a free-body
S
diagram because ma
is not a force.
(c) Incorrect free-body diagram
RIGHT!
You can safely draw
the acceleration
vector to one side
of the diagram.
(c) Incorrect free-body diagram
y
y
The following
problem-solving strategy is very similar to Problem-Solving
y
Strategy 5.1 for equilibrium problems in Section 5.1. Study it carefully, watch
how we apply
x it in our examples, and use it when you tackle the end-of-chapter
problems. You can use this strategy to solve any dynamics problem.
ma
WRONG
S
S
This vector
doesn’t
belong
in free-body diagrams Remember that the quantity ma
caUtion ma doesn’t
belong
in
a
free-body
w
is the result of forces acting on a body,
not a force itself. When you draw the free-body
S
diagram because ma
S
diagram for an accelerating
body
(like
the
fruit in Fig. 5.6a), never include the “ma force”
is not a force.
because there is no such force (Fig. 5.6c). Review Section 4.3 if you’re not clear on this
S
point. Sometimes we draw the acceleration vector a alongside a free-body diagram, as
in Fig. 5.6b. But we never draw the acceleration vector with its tail touching the body
(a position reserved exclusively for forces that act on the body). ❙
WRONG
This vector doesn’t
belong in a free-body
S
diagram because ma
is not a force.
Strategy
5.2 newton’S
Second law: dynamicS oF ParticleS
identiFy the relevant concepts: You have to use Newton’s second
law, Eqs. (5.2), for any problem that involves forces acting on an
accelerating body.
Identify the target variable—usually an acceleration or a force.
If the target variable is something else, you’ll need to select
another concept to use. For example, suppose the target variable
is how fast a sled is moving when it reaches the bottom of a hill.
Newton’s second law will let you find the sled’s acceleration; you’ll
then use the constant-acceleration relationships from Section 2.4
to find velocity from acceleration.
set UP the problem by using the following steps:
1. Draw a simple sketch of the situation that shows each moving
body. For each body, draw a free-body diagram that shows all
the forces acting on the body. [The sums in Eqs. (5.2) include
the forces that act on the body, not the forces that it exerts on
anything else.] Make sure you can answer the question “What
other body is applying this force?” for each force in your
S
diagram. Never include the quantity ma in your free-body
diagram; it’s not a force!
2. Label each force with an algebraic symbol for the force’s
magnitude. Usually, one of the forces will be the body’s
weight; it’s usually best to label this as w = mg.
3. Choose your x- and y-coordinate axes for each body, and show
them in its free-body diagram. Indicate the positive direction
for each axis. If you know the direction of the acceleration, it
usually simplifies things to take one positive axis along that
direction. If your problem involves two or more bodies that
accelerate in different directions, you can use a different set of
axes for each body.
4. In addition to Newton’s second law, gF = ma, identify any other
equations you might need. For example, you might need one or
more of the equations for motion with constant acceleration. If
more than one body is involved, there may be relationships
among their motions; for example, they may be connected by a
rope. Express any such relationships as equations relating the
accelerations of the various bodies.
S
S
execUte the solution as follows:
1. For each body, determine the components of the forces along
each of the body’s coordinate axes. When you represent a force
in terms of its components, draw a wiggly line through the
original force vector to remind you not to include it twice.
2. List all of the known and unknown quantities. In your list,
identify the target variable or variables.
3. For each body, write a separate equation for each component
of Newton’s second law, as in Eqs. (5.2). Write any additional
equations that you identified in step 4 of “Set Up.” (You need
as many equations as there are target variables.)
4. Do the easy part—the math! Solve the equations to find the
target variable(s).
evaLUate your answer: Does your answer have the correct units?
(When appropriate, use the conversion 1 N = 1 kg # m>s2.) Does
it have the correct algebraic sign? When possible, consider particular values or extreme cases of quantities and compare the
results with your intuitive expectations. Ask, “Does this result
make sense?”
5.2 Using Newton’s Second Law: Dynamics of particles
Solution
ExamplE 5.6
straight-linE motion with a constant forcE
An iceboat is at rest on a frictionless horizontal surface (Fig. 5.7a).
Due to the blowing wind, 4.0 s after the iceboat is released, it is
moving to the right at 6.0 m>s (about 22 km>h, or 13 mi>h). What
constant horizontal force FW does the wind exert on the iceboat?
The combined mass of iceboat and rider is 200 kg.
Solution
identify and Set up: Our target variable is one of the forces
1FW2 acting on the accelerating iceboat, so we need to use
Newton’s second law. The forces acting on the iceboat and rider
(considered as a unit) are the weight w, the normal force n exerted
by the surface, and the horizontal force FW. Figure 5.7b shows
the free-body diagram. The net force and hence the acceleration
are to the right, so we chose the positive x-axis in this direction.
The acceleration isn’t given; we’ll need to find it. Since the wind
is assumed to exert a constant force, the resulting acceleration is
constant and we can use one of the constant-acceleration formulas
from Section 2.4.
The iceboat starts at rest (its initial x-velocity is v0x = 0) and it
attains an x-velocity vx = 6.0 m>s after an elapsed time t = 4.0 s.
To relate the x-acceleration ax to these quantities we use Eq. (2.8),
vx = v0x + ax t. There is no vertical acceleration, so we expect
that the normal force on the iceboat is equal in magnitude to the
iceboat’s weight.
execute: The known quantities are the mass m = 200 kg, the
initial and final x-velocities v0x = 0 and vx = 6.0 m>s, and the
elapsed time t = 4.0 s. There are three unknown quantities: the acceleration ax, the normal force n, and the horizontal force FW.
Hence we need three equations.
The first two equations are the x- and y-equations for Newton’s
second law, Eqs. (5.2). The force FW is in the positive x-direction,
while the forces n and w = mg are in the positive and negative
y-directions, respectively. Hence we have
gFx = FW = max
gFy = n + 1- mg2 = 0
so
n = mg
The third equation is Eq. (2.8) for constant acceleration:
5.7 Our sketches for this problem.
(a) Iceboat and rider on frictionless ice
137
(b) Free-body diagram
for iceboat and rider
vx = v0x + axt
To find FW, we first solve this third equation for ax and then
substitute the result into the gFx equation:
ax =
vx - v0x
6.0 m>s - 0 m>s
=
= 1.5 m>s2
t
4.0 s
FW = max = 1200 kg211.5 m>s22 = 300 kg # m>s2
Since 1 kg # m>s2 = 1 N, the final answer is
FW = 300 N 1about 67 lb2
evaluate: Our answers for FW and n have the correct units for
a force, and (as expected) the magnitude n of the normal force is
equal to mg. Does it seem reasonable that the force FW is substantially less than mg?
Solution
ExamplE 5.7
straight-linE motion with friction
Suppose a constant horizontal friction force with magnitude 100 N
opposes the motion of the iceboat in Example 5.6. In this case,
what constant force FW must the wind exert on the iceboat to
cause the same constant x-acceleration ax = 1.5 m>s2?
5.8 Our
free-body diagram for the iceboat and rider with friction
S
force f opposing the motion.
Solution
identify and Set up: Again the target variable is FW. We are
given the x-acceleration, so to find FW all we need is Newton’s
second law. Figure 5.8 shows our new free-body diagram. The
only difference
from Fig. 5.7b is the addition of the friction
S
force ƒ, which points in the negative x-direction (opposite the
motion). Because the wind must now overcome the friction force
to yield the same acceleration as in Example 5.6, we expect our
answer for FW to be greater than the 300 N we found there.
Continued
138
Chapter 5 applying Newton’s Laws
execute: Two forces now have x-components: the force of the
wind (x-component + FW) and the friction force (x-component -f ).
The x-component of Newton’s second law gives
gFx = FW + 1- f 2 = max
evaluate: The required value of FW is 100 N greater than in
Example 5.6 because the wind must now push against an additional 100-N friction force.
FW = max + f = 1200 kg211.5 m>s22 + 1100 N2 = 400 N
Solution
ExamplE 5.8
tEnsion in an ElEvator cablE
An elevator and its load have a combined mass of 800 kg (Fig. 5.9a).
The elevator is initially moving downward at 10.0 m>s; it slows to
a stop with constant acceleration in a distance of 25.0 m. What is
the tension T in the supporting cable while the elevator is being
brought to rest?
Solution
identify and Set up: The target variable is the tension T, which
we’ll find by using Newton’s second law. As in Example 5.6, we’ll
use a constant-acceleration formula to determine the acceleration.
Our free-body diagram (Fig. 5.9b) shows two forces acting on the
elevator: its weight w and the tension force T of the cable. The elevator is moving downward with decreasing speed, so its acceleration is upward; we chose the positive y-axis to be upward.
execute: First let’s write out Newton’s second law. The tension
force acts upward and the weight acts downward, so
gFy = T + 1-w2 = may
We solve for the target variable T:
T = w + may = mg + may = m1g + ay2
5.9 Our sketches for this problem.
(a) Descending elevator
The elevator is moving in the negative y-direction, so both its
initial y-velocity v0y and its y-displacement y - y0 are negative:
v0y = -10.0 m>s and y - y0 = - 25.0 m. The final y-velocity is
vy = 0. To find the y-acceleration ay from this information, we’ll
use Eq. (2.13) in the form v y2 = v 0y 2 + 2ay1y - y02. Once we
have ay , we’ll substitute it into the y-component of Newton’s second law from Eqs. (5.2) and solve for T. The net force must be upward to give an upward acceleration, so we expect T to be greater
than the weight w = mg = 1800 kg219.80 m>s22 = 7840 N.
(b) Free-body diagram
for elevator
To determine ay , we rewrite the constant-acceleration equation
v y2 = v 0y 2 + 2ay1y - y02:
ay =
v y2 - v 0y 2
21y - y02
=
1022 - 1-10.0 m>s22
21- 25.0 m2
= + 2.00 m>s2
The acceleration is upward (positive), just as it should be.
Now we can substitute the acceleration into the equation for
the tension:
Moving down with
decreasing speed
evaluate: The tension is greater than the weight, as expected. Can
you see that we would get the same answers for ay and T if the elevator were moving upward and gaining speed at a rate of 2.00 m>s2?
Solution
ExamplE 5.9
T = m1g + ay2 = 1800 kg219.80 m>s2 + 2.00 m>s22 = 9440 N
apparEnt wEight in an accElErating ElEvator
A 50.0-kg woman stands on a bathroom scale while riding in the
elevator in Example 5.8. What is the reading on the scale?
5.10 Our sketches for this problem.
(a) Woman in a
descending elevator
(b) Free-body diagram
for woman
Solution
identify and Set up: The scale (Fig. 5.10a) reads the magni-
tude of the downward force exerted by the woman on the scale.
By Newton’s third law, this equals the magnitude of the upward
normal force exerted by the scale on the woman. Hence our target
variable is the magnitude n of the normal force. We’ll find n by
applying Newton’s second law to the woman. We already know
her acceleration; it’s the same as the acceleration of the elevator,
which we calculated in Example 5.8.
Figure 5.10b shows our free-body diagram for the woman.
The forces acting on her are the normal force n exerted by the
Moving down with
decreasing speed
5.2 Using Newton’s Second Law: Dynamics of Particles
scale and her weight w = mg = 150.0 kg219.80 m>s22 = 490 N.
(The tension force, which played a major role in Example 5.8,
doesn’t appear here because it doesn’t act on the woman.) From
Example 5.8, the y-acceleration of the elevator and of the woman is
ay = + 2.00 m>s2. As in Example 5.8, the upward force on the body
accelerating upward (in this case, the normal force on the woman)
will have to be greater than the body’s weight to produce the
upward acceleration.
execUte: Newton’s second law gives
gFy = n + 1- mg2 = may
n = mg + may = m1g + ay2
= 150.0 kg219.80 m>s2 + 2.00 m>s22 = 590 N
evaLUate: Our answer for n means that while the elevator is stopping, the scale pushes up on the woman with a force of 590 N. By
Newton’s third law, she pushes down on the scale with the same
139
force. So the scale reads 590 N, which is 100 N more than her actual
weight. The scale reading is called the passenger’s apparent weight.
The woman feels the floor pushing up harder on her feet than when
the elevator is stationary or moving with constant velocity.
What would the woman feel if the elevator were accelerating
downward, so that ay = -2.00 m>s2? This would be the case if
the elevator were moving upward with decreasing speed or moving downward with increasing speed. To find the answer for this
situation, we just insert the new value of ay in our equation for n:
n = m1g + ay2 = 150.0 kg239.80 m>s2 + 1- 2.00 m>s224
= 390 N
Now the woman would feel as though she weighs only 390 N, or
100 N less than her actual weight w.
You can feel these effects yourself; try taking a few steps in an
elevator that is coming to a stop after descending (when your apparent weight is greater than w) or coming to a stop after ascending (when your apparent weight is less than w).
apparent weight and apparent weightlessness
Let’s generalize the result of Example 5.9. When a passenger with mass m rides
in an elevator with y-acceleration ay , a scale shows the passenger’s apparent
weight to be
n = m1g + ay2
5.11 Astronauts in orbit feel “weightless”
because they have the same acceleration
as their spacecraft. They are not outside
the pull of the earth’s gravity. (We’ll discuss
the motions of orbiting bodies in detail in
Chapter 12.)
When the elevator is accelerating upward, ay is positive and n is greater than the
passenger’s weight w = mg. When the elevator is accelerating downward, ay is
negative and n is less than the weight. If the passenger doesn’t know the elevator
is accelerating, she may feel as though her weight is changing; indeed, this is just
what the scale shows.
The extreme case occurs when the elevator has a downward acceleration
ay = -g—that is, when it is in free fall. In that case n = 0 and the passenger
seems to be weightless. Similarly, an astronaut orbiting the earth with a spacecraft experiences apparent weightlessness (Fig. 5.11). In each case, the person
is not truly weightless because a gravitational force still acts. But the person’s
sensations in this free-fall condition are exactly the same as though the person
were in outer space with no gravitational force at all. In both cases the person and
the vehicle (elevator or spacecraft) fall together with the same acceleration g, so
nothing pushes the person against the floor or walls of the vehicle.
Solution
examPle 5.10
acceleration down a hill
A toboggan loaded with students (total weight w) slides down a
snow-covered hill that slopes at a constant angle a. The toboggan is
well waxed, so there is virtually no friction. What is its acceleration?
5.12 Our sketches for this problem.
(a) The situation
(b) Free-body diagram for toboggan
soLUtion
identiFy and set UP: Our target variable is the acceleration,
which we’ll find by using Newton’s second law. There is no friction, so only two forces act on the toboggan: its weight w and the
normal force n exerted by the hill.
Figure 5.12 shows our sketch and free-body diagram. We take
axes parallel and perpendicular to the surface of the hill, so that
the acceleration (which is parallel to the hill) is along the positive
x-direction.
Continued
140
ChaPtEr 5 applying Newton’s Laws
execUte: The normal force has only a y-component, but the
evaLUate: Notice that the normal force n is not equal to the
weight has both x- and y-components: wx = w sin a and wy =
- w cos a. (In Example 5.4 we had wx = - w sin a. The difference
is that the positive x-axis was uphill in Example 5.4 but is downhill in Fig. 5.12b.) The wiggly line in Fig. 5.12b reminds us that
we have resolved the weight into its components. The acceleration
is purely in the + x@direction, so ay = 0. Newton’s second law in
component form from Eqs. (5.2) then tells us that
toboggan’s weight (compare Example 5.4). Notice also that the
mass m does not appear in our result for the acceleration. That’s
because the downhill force on the toboggan (a component of the
weight) is proportional to m, so the mass cancels out when we
use gFx = max to calculate ax. Hence any toboggan, regardless of its mass, slides down a frictionless hill with acceleration
g sin a.
If the plane is horizontal, a = 0 and ax = 0 (the toboggan
does not accelerate); if the plane is vertical, a = 90° and ax = g
(the toboggan is in free fall).
gFx = w sin a = max
gFy = n - w cos a = may = 0
Since w = mg, the x-component equation gives mg sin a = max, or
ax = g sin a
Note that we didn’t need the y-component equation to find the acceleration. That’s part of the beauty of choosing the x-axis to lie
along the acceleration direction! The y-equation tells us the magnitude of the normal force exerted by the hill on the toboggan:
n = w cos a = mg cos a
caUtion common free-body diagram errors Figure 5.13 shows
both the correct way (Fig. 5.13a) and a common incorrect way
(Fig. 5.13b) to draw the free-body diagram for the toboggan.
The diagram in Fig. 5.13b is wrong for two reasons: The normal
force must be drawn perpendicular to the surface (remember,
“normal” means perpendicular), and there’s no such thing as the
S
“ma force.” ❙
5.13 Correct and incorrect free-body diagrams for a toboggan on a frictionless hill.
(a) Correct free-body diagram for the sled
RIGHT!
Normal force is
perpendicular
to the surface.
It’s OK to draw the
acceleration vector
adjacent to (but not
touching) the body.
RIGHT!
(b) Incorrect free-body diagram for the sled
WRONG
Normal force is not
vertical because the
surface (which is
along the x-axis)
is inclined.
The quantity ma is
not a force.
WRONG
Solution
examPle 5.11 two bodieS with the Same acceleration
You push a 1.00-kg food tray through the cafeteria line with a
constant 9.0-N force. The tray pushes a 0.50-kg milk carton
(Fig. 5.14a). The tray and carton slide on a horizontal surface so
greasy that friction can be ignored. Find the acceleration of the tray
and carton and the horizontal force that the tray exerts on the carton.
soLUtion
identiFy and set UP: Our two target variables are the acceleration of the tray–carton system and the force of the tray on the carton. We’ll use Newton’s second law to get two equations, one for
each target variable. We set up and solve the problem in two ways.
Method 1: We treat the carton (mass m C) and tray (mass m T) as
separate bodies, each with its own free-body diagram (Figs. 5.14b
and 5.14c). The force F that you exert on the tray doesn’t appear in
the free-body diagram for the carton, which is accelerated by the
force (of magnitude FT on C) exerted on it by the tray. By Newton’s
third law, the carton exerts a force of equal magnitude on the tray:
FC on T = FT on C. We take the acceleration to be in the positive
x-direction; both the tray and milk carton move with the same
x-acceleration ax .
Method 2: We treat the tray and milk carton as a composite
body of mass m = m T + m C = 1.50 kg (Fig. 5.14d). The only
horizontal force acting on this body is the force F that you exert.
The forces FT on C and FC on T don’t come into play because
they’re internal to this composite body, and Newton’s second
law tells us that only external forces affect a body’s acceleration
(see Section 4.3). To find the magnitude FT on C we’ll again apply
Newton’s second law to the carton, as in Method 1.
execUte: Method 1: The x-component equations of Newton’s
second law are
Tray:
Carton:
gFx = F - FC on T = F - FT on C = m Tax
gFx = FT on C = m Cax
These are two simultaneous equations for the two target variables
ax and FT on C. (Two equations are all we need, which means that
the y-components don’t play a role in this example.) An easy way
to solve the two equations for ax is to add them; this eliminates
FT on C, giving
F = m Tax + m Cax = 1m T + m C2ax
5.2 Using Newton’s Second Law: Dynamics of Particles
141
5.14 Pushing a food tray and milk carton in the cafeteria line.
(a) A milk carton and a food tray
(b) Free-body diagram
for milk carton
(c) Free-body diagram
for food tray
(d) Free-body diagram for
carton and tray as a composite body
y
y
mC = 0.50 kg
F = 9.0 N
ax
x
FT on C
y
ax
nC
F
x
n
nT
FC on T =
FT on C
ax
x
F
wC
wT
mT = 1.00 kg
w
We solve this equation for ax :
ax =
F
9.0 N
=
= 6.0 m>s2 = 0.61g
mT + mC
1.00 kg + 0.50 kg
Substituting this value into the carton equation gives
FT on C = m Cax = 10.50 kg216.0 m>s22 = 3.0 N
Method 2: The x-component of Newton’s second law for the
composite body of mass m is
gFx = F = max
The acceleration of this composite body is
ax =
FT on C = m Cax = 10.50 kg216.0 m>s22 = 3.0 N
evaLUate: The answers are the same with both methods. To check
the answers, note that there are different forces on the two sides
of the tray: F = 9.0 N on the right and FC on T = 3.0 N on the
left. The net horizontal force on the tray is F - FC on T = 6.0 N,
exactly enough to accelerate a 1.00-kg tray at 6.0 m>s2.
Treating two bodies as a single, composite body works only
if the two bodies have the same magnitude and direction of acceleration. If the accelerations are different we must treat the two
bodies separately, as in the next example.
Solution
examPle 5.12
F
9.0 N
= 6.0 m>s2
=
m
1.50 kg
Then, looking at the milk carton by itself, we see that to give it an
acceleration of 6.0 m>s2 requires that the tray exert a force
two bodieS with the Same magnitude oF acceleration
Figure 5.15a shows an air-track glider with mass m 1 moving
on a level, frictionless air track in the physics lab. The glider is
connected to a lab weight with mass m 2 by a light, flexible, nonstretching string that passes over a stationary, frictionless pulley.
Find the acceleration of each body and the tension in the string.
5.15 Our sketches for this problem.
(a) Apparatus
(b) Free-body
diagram for glider
(c) Free-body
diagram for weight
m1
soLUtion
identiFy and set UP: The glider and weight are accelerating,
so again we must use Newton’s second law. Our three target variables are the tension T in the string and the accelerations of the
two bodies.
The two bodies move in different directions—one horizontal,
one vertical—so we can’t consider them to be a single unit as we
did the bodies in Example 5.11. Figures 5.15b and 5.15c show our
free-body diagrams and coordinate systems. It’s convenient to
have both bodies accelerate in the positive axis directions, so we
chose the positive y-direction for the lab weight to be downward.
We consider the string to be massless and to slide over the
pulley without friction, so the tension T in the string is the same
throughout and it applies a force of the same magnitude T to each
m2
body. (You may want to review Conceptual Example 4.10, in
which we discussed the tension force exerted by a massless rope.)
The weights are m 1g and m 2g.
Continued
142
ChaPtEr 5 applying Newton’s Laws
While the directions of the two accelerations are different,
their magnitudes are the same. (That’s because the string doesn’t
stretch, so the two bodies must move equal distances in equal
times and their speeds at any instant must be equal. When the
speeds change, they change at the same rate, so the accelerations
of the two bodies must have the same magnitude a.) We can express
this relationship as a1x = a2y = a, which means that we have only
two target variables: a and the tension T.
What results do we expect? If m 1 = 0 (or, approximately, for
m 1 much less than m 2) the lab weight will fall freely with acceleration g, and the tension in the string will be zero. For m 2 = 0
(or, approximately, for m 2 much less than m 1) we expect zero
acceleration and zero tension.
We add the two equations to eliminate T, giving
execUte: Newton’s second law gives
expect; the string tension keeps the lab weight from falling freely.
The tension T is not equal to the weight m 2g of the lab weight, but
is less by a factor of m 1>1m 1 + m 22. If T were equal to m 2g, then
the lab weight would be in equilibrium, and it isn’t.
As predicted, the acceleration is equal to g for m 1 = 0 and
equal to zero for m 2 = 0, and T = 0 for either m 1 = 0 or m 2 = 0.
Glider: gFx = T = m 1a1x = m 1a
Glider: gFy = n + 1- m 1g2 = m 1a1y = 0
Lab weight: gFy = m 2g + 1- T2 = m 2a2y = m 2a
(There are no forces on the lab weight in the x-direction.) In these
equations we’ve used a1y = 0 (the glider doesn’t accelerate vertically) and a1x = a2y = a.
The x-equation for the glider and the equation for the lab
weight give us two simultaneous equations for T and a:
Glider: T = m 1a
Lab weight: m 2g - T = m 2a
PhET: Lunar Lander
m 2g = m 1a + m 2a = 1m 1 + m 22a
and so the magnitude of each body’s acceleration is
a =
m2
g
m1 + m2
Substituting this back into the glider equation T = m 1a, we get
T =
m 1m 2
g
m1 + m2
evaLUate: The acceleration is in general less than g, as you might
caUtion tension and weight may not be equal It’s a common
mistake to assume that if an object is attached to a vertical string,
the string tension must be equal to the object’s weight. That was
the case in Example 5.5, where the acceleration was zero, but it’s
not the case in this example! The only safe approach is always to
treat the tension as a variable, as we did here. ❙
test yoUr Understanding oF section 5.2 Suppose you hold the glider
in Example 5.12 so that it and the weight are initially at rest. You give the glider a push
to the left in Fig. 5.15a and then release it. The string remains taut as the glider moves to
the left, comes instantaneously to rest, then moves to the right. At the instant the glider
has zero velocity, what is the tension in the string? (i) Greater than in Example 5.12;
(ii) the same as in Example 5.12; (iii) less than in Example 5.12 but greater than zero;
(iv) zero. ❙
5.3 Friction Forces
5.16 There is friction between the feet of
this caterpillar (the larval stage of a butterfly of the family Papilionidae) and the
surfaces over which it walks. Without
friction, the caterpillar could not move
forward or climb over obstacles.
We’ve seen several problems in which a body rests or slides on a surface that
exerts forces on the body. Whenever two bodies interact by direct contact (touching) of their surfaces, we describe the interaction in terms of contact forces. The
normal force is one example of a contact force; in this section we’ll look in detail
at another contact force, the force of friction.
Friction is important in many aspects of everyday life. The oil in a car engine
minimizes friction between moving parts, but without friction between the tires
and the road we couldn’t drive or turn the car. Air drag—the friction force exerted by the air on a body moving through it—decreases automotive fuel economy but makes parachutes work. Without friction, nails would pull out and most
forms of animal locomotion would be impossible (Fig. 5.16).
Kinetic and static Friction
When you try to slide a heavy box of books across the floor, the box doesn’t
move at all unless you push with a certain minimum force. Once the box starts
moving, you can usually keep it moving with less force than you needed to get it
started. If you take some of the books out, you need less force to get it started or
keep it moving. What can we say in general about this behavior?
143
5.3 Friction Forces
First, when a body rests or slides on a surface, we can think of the surface as
exerting a single contact force on the body, with force components perpendicular
and parallel to the surface (Fig. 5.17). The perpendicular component vector is the
S
normal force, denoted by n. The component vector parallel
to the surface (and
S
S
perpendicular
to
n
)
is
the
friction
force,
denoted
by
f
.
If
the
surface is frictionS
less, then f is zero but there is still a normal force. (Frictionless surfaces are
an unattainable idealization, like a massless rope. But we can approximate a
surface as frictionless if the effects of friction are negligibly small.) The direction of the friction force is always such as to oppose relative motion of the two
surfaces.
The kind of frictionSthat acts when a body slides over a surface is called a
kinetic friction force fk . The adjective “kinetic” and the subscript “k” remind
us that the two surfaces are moving relative to each other. The magnitude of
the kinetic friction force usually increases when the normal force increases.
This is why it takes more force to slide a full box of books across the floor than
an empty one. Automotive brakes use the same principle: The harder the brake
pads are squeezed against the rotating brake discs, the greater the braking
effect. In many cases the magnitude of the kinetic friction force fk is found
experimentally to be approximately proportional to the magnitude n of the
normal force:
Magnitude of kinetic
friction force
f k = mk n
Coefficient of kinetic friction
Magnitude of normal force
5.17 When a block is pushed or pulled
over a surface, the surface exerts a contact
force on it.
The friction and normal force ...
... are components of
a single contact force.
Contact force
Normal-force
component n
Friction-force
component f
Push or pull
Weight mg
5.18 A microscopic view of the friction
and normal forces.
Block
(5.3)
Floor
Magnified view
Here mk (pronounced “mu-sub-k”) is a constant called the coefficient of kinetic
friction. The more slippery the surface, the smaller this coefficient. Because it is
a quotient of two force magnitudes, mk is a pure number without units.
are always perpendicular Remember that Eq. (5.3)
caUtion Friction and normal forces
S
S
is not a vector equation because fk and n are always perpendicular. Rather, it is a scalar
relationship between the magnitudes of the two forces. ❙
Equation (5.3) is only an approximate representation of a complex phenomenon.
On a microscopic level, friction and normal forces result from the intermolecular forces (electrical in nature) between two rough surfaces at points where they
come into contact (Fig. 5.18). As a box slides over the floor, bonds between the
two surfaces form and break, and the total number of such bonds varies. Hence
the kinetic friction force is not perfectly constant. Smoothing the surfaces can
actually increase friction, since more molecules can interact and bond; bringing
two smooth surfaces of the same metal together can cause a “cold weld.”
Lubricating oils work because an oil film between two surfaces (such as the
pistons and cylinder walls in a car engine) prevents them from coming into
actual contact.
Table 5.1 lists some representative values of mk. Although these values are
given with two significant figures, they are only approximate, since friction
forces can also depend on the speed of the body relative to the surface. For now
we’ll ignore this effect and assume that mk and fk are independent of speed, in
order to concentrate on the simplest cases. Table 5.1 also lists coefficients of
static friction; we’ll define these shortly.
Friction forces may also act when there is no relative motion. If you try to
slide a box across the floor, the box may not move at all because the floor exerts
an equal
and opposite friction force on the box. This is called a static friction
S
force fs .
The friction and normal forces result from
interactions between molecules in the block and
in the floor where the two rough surfaces touch.
approximate
Table 5.1 coefficients of Friction
Materials
Coefficient Coefficient
of Static
of Kinetic
Friction,
Friction,
Ms
Mk
Steel on steel
0.74
0.57
Aluminum on steel
0.61
0.47
Copper on steel
0.53
0.36
Brass on steel
0.51
0.44
Zinc on cast iron
0.85
0.21
Copper on cast iron
1.05
0.29
Glass on glass
0.94
0.40
Copper on glass
0.68
0.53
Teflon on Teflon
0.04
0.04
Teflon on steel
0.04
0.04
Rubber on concrete
(dry)
1.0
0.8
Rubber on concrete
(wet)
0.30
0.25
144
Chapter 5 applying Newton’s Laws
5.19 When there is no relative motion, the magnitude of the static friction force fs is less than
or equal to ms n. When there is relative motion, the magnitude of the kinetic friction force fk
equals mkn.
(a)
n
(b)
(c)
n
T
fs
w
1 No applied force,
(d)
n
T
fs
w
T
fk
w
w
2 Weak applied force,
3 Stronger applied force,
box remains at rest.
Static friction:
fs 6 ms n
box just about to slide.
Static friction:
fs = m s n
box at rest.
No friction:
fs = 0
n
4 Box sliding at
constant speed.
Kinetic friction:
fk = mk n
f
(e)
1 fs2max
fk
5 This graph shows the friction
8 The kinetic friction force
force magnitude f as a function
of the pulling force magnitude T.
varies somewhat as
intermolecular bonds
form and break.
T
O
6 Box at rest; static friction
7 Box moving; kinetic friction
equals applied force.
is essentially constant.
S
application Static friction and
Windshield Wipers The squeak of
windshield wipers on dry glass is a stick-slip
phenomenon. The moving wiper blade sticks
to the glass momentarily, then slides when
the force applied to the blade by the wiper
motor overcomes the maximum force of static
friction. When the glass is wet from rain or
windshield cleaning solution, friction is
reduced and the wiper blade doesn’t stick.
In Fig. 5.19a, the box is at rest, in equilibrium, under the action of its weight w
S
and the upward normal force n. The normal force is equal in magnitude to
the weight 1n = w2 and is exerted on the box by the floor. Now we tie a rope to
the box (Fig. 5.19b) and gradually increase the tension T in the rope. At first the
box remains at rest because the force of static friction fs also increases and stays
equal in magnitude to T.
At some point T becomes greater than the maximum static friction force fs the
surface can exert. Then the box “breaks loose” and starts to slide. Figure 5.19c
shows the forces when T is at this critical value. For a given pair of surfaces the
maximum value of fs depends on the normal force. Experiment shows that in
many cases this maximum value, called 1fs2max , is approximately proportional
to n; we call the proportionality factor ms the coefficient of static friction.
Table 5.1 lists some representative values of ms . In a particular situation, the
actual force of static friction can have any magnitude between zero (when
there is no other force parallel to the surface) and a maximum value given
by msn:
Magnitude of static
friction force
fs … 1 fs2max = msn
Maximum static friction force
Coefficient of static friction
(5.4)
Magnitude of normal force
Like Eq. (5.3), this is a relationship between magnitudes, not a vector relationship. The equality sign holds only when the applied force T has reached the
critical value at which motion is about to start (Fig. 5.19c). When T is less than
this value (Fig. 5.19b), the Sinequality sign holds. In that case we have to use the
equilibrium conditions 1gF = 02 to find fs . If there is no applied force 1T = 02
as in Fig. 5.19a, then there is no static friction force either 1fs = 02.
As soon as the box starts to slide (Fig. 5.19d), the friction force usually decreases (Fig. 5.19e); it’s easier to keep the box moving than to start it moving.
5.3 Friction Forces
Hence the coefficient of kinetic friction is usually less than the coefficient of
static friction for any given pair of surfaces, as Table 5.1 shows.
In some situations the surfaces will alternately stick (static friction) and
slip (kinetic friction). This is what causes the horrible sound made by chalk
held at the wrong angle on a blackboard and the shriek of tires sliding on asphalt pavement. A more positive example is the motion of a violin bow against
the string.
In the linear air tracks used in physics laboratories, gliders move with very
little friction because they are supported on a layer of air. The friction force is
velocity dependent, but at typical speeds the effective coefficient of friction is of
the order of 0.001.
PhET: Friction
PhET: The Ramp
Friction in horizontal motion
You want to move a 500-N crate across a level floor. To start the
crate moving, you have to pull with a 230-N horizontal force.
Once the crate starts to move, you can keep it moving at constant
velocity with only 200 N. What are the coefficients of static and
kinetic friction?
Solution
identify and Set up: The crate is in equilibrium both when it
is at rest and when it is moving at constant velocity, so we use
Newton’s first law, as expressed by Eqs. (5.1). We use Eqs. (5.3)
and (5.4) to find the target variables ms and mk.
Figures 5.20a and 5.20b show our sketch and free-body diagram for the instant just before the crate starts to move, when the
static friction force has its maximum possible value 1fs2max = msn.
5.20 Our sketches for this problem.
(a) Pulling a crate
PhET: Forces in 1 Dimension
Solution
ExamplE 5.13
145
(b) Free-body diagram
for crate just before it
starts to move
(c) Free-body diagram
for crate moving at
constant speed
Once the crate is moving, the friction force changes to its kinetic
form (Fig. 5.20c). In both situations, four forces act on the crate:
the downward weight (magnitude w = 500 N), the upward normal
force (magnitude n) exerted by the floor, a tension force (magnitude T) to the right exerted by the rope, and a friction force to the
left exerted by the floor. Because the rope in Fig. 5.20a is in equilibrium, the tension is the same at both ends. Hence the tension
force that the rope exerts on the crate has the same magnitude as
the force you exert on the rope. Since it’s easier to keep the crate
moving than to start it moving, we expect that mk 6 ms.
execute: Just before the crate starts to move (Fig. 5.20b), we
have from Eqs. (5.1)
gFx = T + 1-1 fs2max2 = 0 so
gFy = n + 1- w2 = 0
so
1 fs2max = T = 230 N
n = w = 500 N
Now we solve Eq. (5.4), 1fs2max = msn, for the value of ms:
ms =
1 fs2max
n
=
230 N
= 0.46
500 N
After the crate starts to move (Fig. 5.20c) we have
gFx = T + 1-fk2 = 0 so
gFy = n + 1- w2 = 0
fk = T = 200 N
so n = w = 500 N
Using fk = mkn from Eq. (5.3), we find
mk =
fk
200 N
=
= 0.40
n
500 N
evaluate: As expected, the coefficient of kinetic friction is less
than the coefficient of static friction.
Solution
ExamplE 5.14
Static Friction can bE lESS than thE maximum
In Example 5.13, what is the friction force if the crate is at rest on
the surface and a horizontal force of 50 N is applied to it?
diagram is the same as in Fig. 5.20b, but with 1 fs2max replaced by
fs and T = 230 N replaced by T = 50 N.
execute: From the equilibrium conditions, Eqs. (5.1), we have
Solution
identify and Set up: The applied force is less than the maxi-
mum force of static friction, 1 fs2max = 230 N. Hence the crate
remains at rest and the net force acting on it is zero. The target
variable is the magnitude fs of the friction force. The free-body
gFx = T + 1-fs2 = 0
so
fs = T = 50 N
evaluate: The friction force can prevent motion for any horizontal applied force up to 1 fs2max = msn = 230 N. Below that value,
fs has the same magnitude as the applied force.
146
ChaPtEr 5 applying Newton’s Laws
Solution
examPle 5.15
minimizing kinetic Friction
In Example 5.13, suppose you move the crate by pulling upward
on the rope at an angle of 30° above the horizontal. How hard
must you pull to keep it moving with constant velocity? Assume
that mk = 0.40.
force n is not equal in magnitude to the crate’s weight. The force
exerted by the rope has a vertical component that tends to lift the
crate off the floor; this reduces n and so reduces fk.
execUte: From the equilibrium conditions and Eq. (5.3), fk =
mkn, we have
soLUtion
identiFy and set UP: The crate is in equilibrium because its ve-
locity is constant, so we again apply Newton’s first law. Since the
crate is in motion, the floor exerts a kinetic friction force. The target variable is the magnitude T of the tension force.
Figure 5.21 shows our sketch and free-body diagram. The kinetic friction force fk is still equal to mkn, but now the normal
5.21 Our sketches for this problem.
(b) Free-body diagram for moving crate
(a) Pulling a crate at an angle
gFx = T cos 30° + 1-fk2 = 0
so
T cos 30° = mkn
gFy = T sin 30° + n + 1-w2 = 0 so n = w - T sin 30°
These are two equations for the two unknown quantities T and n.
One way to find T is to substitute the expression for n in the second equation into the first equation and then solve the resulting
equation for T:
T cos 30° = mk1w - T sin 30°2
mkw
T =
= 188 N
cos 30° + mk sin 30°
We can substitute this result into either of the original equations to
obtain n. If we use the second equation, we get
n = w - T sin 30° = 1500 N2 - 1188 N2 sin 30° = 406 N
evaLUate: As expected, the normal force is less than the 500-N
weight of the box. It turns out that the tension required to keep the
crate moving at constant speed is a little less than the 200-N force
needed when you pulled horizontally in Example 5.13. Can you
find an angle where the required pull is minimum?
Solution
examPle 5.16
toboggan ride with Friction i
Let’s go back to the toboggan we studied in Example 5.10. The
wax has worn off, so there is now a nonzero coefficient of kinetic
friction mk. The slope has just the right angle to make the toboggan
slide with constant velocity. Find this angle in terms of w and mk.
5.22 Our sketches for this problem.
(a) The situation
(b) Free-body diagram for toboggan
soLUtion
identiFy and set UP: Our target variable is the slope angle a.
The toboggan is in equilibrium because its velocity is constant, so
we use Newton’s first law in the form of Eqs. (5.1).
Three forces act on the toboggan: its weight, the normal force,
and the kinetic friction force. The motion is downhill, so the friction force (which opposes the motion) is directed uphill. Figure 5.22
shows our sketch and free-body diagram (compare Fig. 5.12b in
Example 5.10). From Eq. (5.3), the magnitude of the kinetic friction
force is fk = mkn. We expect that the greater the value of mk, the
steeper will be the required slope.
execUte: The equilibrium conditions are
gFx = w sin a + 1- fk2 = w sin a - mkn = 0
gFy = n + 1-w cos a2 = 0
Rearranging these two equations, we get
mkn = w sin a and n = w cos a
As in Example 5.10, the normal force is not equal to the weight.
We eliminate n by dividing the first of these equations by the
second, with the result
mk =
sin a
= tan a so a = arctan mk
cos a
evaLUate: The weight w doesn’t appear in this expression. Any
toboggan, regardless of its weight, slides down an incline with
constant speed if the coefficient of kinetic friction equals the tangent of the slope angle of the incline. The arctangent function
increases as its argument increases, so it’s indeed true that the slope
angle a increases as mk increases.
5.3 Friction Forces
Solution
examPle 5.17
147
toboggan ride with Friction ii
The same toboggan with the same coefficient of friction as in
Example 5.16 accelerates down a steeper hill. Derive an expression
for the acceleration in terms of g, a, mk, and w.
From the second equation and Eq. (5.3) we get an expression for fk:
n = mg cos a
fk = mkn = mk mg cos a
We substitute this into the x-component equation and solve for ax:
soLUtion
identiFy and set UP: The toboggan is accelerating, so we must
use Newton’s second law as given in Eqs. (5.2). Our target variable
is the downhill acceleration.
Our sketch and free-body diagram (Fig. 5.23) are almost the
same as for Example 5.16. The toboggan’s y-component of acceleration ay is still zero but the x-component ax is not, so we’ve drawn
w sin a, the downhill component of weight, as a longer vector than
the (uphill) friction force.
execUte: It’s convenient to express the weight as w = mg. Then
Newton’s second law in component form says
gFx = mg sin a + 1- fk2 = max
gFy = n + 1-mg cos a2 = 0
5.23 Our sketches for this problem.
(a) The situation
(b) Free-body diagram for toboggan
mg sin a + 1- mk mg cos a2 = max
ax = g1sin a - mk cos a2
evaLUate: As for the frictionless toboggan in Example 5.10, the
acceleration doesn’t depend on the mass m of the toboggan. That’s
because all of the forces that act on the toboggan (weight, normal
force, and kinetic friction force) are proportional to m.
Let’s check some special cases. If the hill is vertical 1a = 90°2
so that sin a = 1 and cos a = 0, we have ax = g (the toboggan
falls freely). For a certain value of a the acceleration is zero; this
happens if
sin a = mk cos a
and
mk = tan a
This agrees with our result for the constant-velocity toboggan in
Example 5.16. If the angle is even smaller, mk cos a is greater than
sin a and ax is negative; if we give the toboggan an initial downhill push to start it moving, it will slow down and stop. Finally,
if the hill is frictionless so that mk = 0, we retrieve the result of
Example 5.10: ax = g sin a.
Notice that we started with a simple problem (Example 5.10)
and extended it to more and more general situations. The general
result we found in this example includes all the previous ones as
special cases. Don’t memorize this result, but do make sure you
understand how we obtained it and what it means.
Suppose instead we give the toboggan an initial push up the hill.
The direction of the kinetic friction force is now reversed, so the
acceleration is different from the downhill value. It turns out that
the expression for ax is the same as for downhill motion except that
the minus sign becomes plus. Can you show this?
rolling Friction
It’s a lot easier to move a loaded filing cabinet across a horizontal floor by
using a cart with wheels than by sliding it. How much easier? We can define a
coefficient of rolling friction mr , which is the horizontal force needed for constant speed on a flat surface divided by the upward normal force exerted by the
surface. Transportation engineers call mr the tractive resistance. Typical values of
mr are 0.002 to 0.003 for steel wheels on steel rails and 0.01 to 0.02 for rubber tires on concrete. These values show one reason railroad trains are generally
much more fuel efficient than highway trucks.
data SpeakS
static Friction
When students were given a problem
about the magnitude fs of the static friction force acting on an object at rest,
more than 36% gave an incorrect answer.
Common errors:
●
Fluid resistance and terminal speed
Sticking your hand out the window of a fast-moving car will convince you of the
existence of fluid resistance, the force that a fluid (a gas or liquid) exerts on a
body moving through it. The moving body exerts a force on the fluid to push it
out of the way. By Newton’s third law, the fluid pushes back on the body with an
equal and opposite force.
The direction of the fluid resistance force acting on a body is always opposite
the direction of the body’s velocity relative to the fluid. The magnitude of the fluid
resistance force usually increases with the speed of the body through the fluid.
●
Assuming that fs is always equal to msn
(coefficient of static friction * normal
force on object). This is the maximum
value of fs ; the actual value can be anything between zero and this maximum.
Forgetting to apply Newton’s first law to
the object at rest. This is the only correct
way to find the value of fs required to
keep the object from accelerating.
148
ChaPtEr 5 applying Newton’s Laws
This is very different from the kinetic friction force between two surfaces in
contact, which we can usually regard as independent of speed. For small objects
moving at very low speeds, the magnitude f of the fluid resistance force is
approximately proportional to the body’s speed v:
5.24 Motion with fluid resistance.
(a) Metal ball falling
through oil
(b) Free-body diagram
for ball in oil
f = kv
(fluid resistance at low speed)
(5.5)
where k is a proportionality constant that depends on the shape and size of the
body and the properties of the fluid. Equation (5.5) is appropriate for dust particles falling in air or a ball bearing falling in oil. For larger objects moving
through air at the speed of a tossed tennis ball or faster, the resisting force is
approximately proportional to v 2 rather than to v. It is then called air drag or
simply drag. Airplanes, falling raindrops, and bicyclists all experience air drag.
In this case we replace Eq. (5.5) by
f
x
w = mg
f = Dv 2
y
(fluid resistance at high speed)
(5.6)
2
Because of the v dependence, air drag increases rapidly with increasing speed.
The air drag on a typical car is negligible at low speeds but comparable to or
greater than rolling resistance at highway speeds. The value of D depends on the
shape and size of the body and on the density of the air. You should verify that
the units of the constant k in Eq. (5.5) are N # s>m or kg>s, and that the units of
the constant D in Eq. (5.6) are N # s2>m2 or kg>m.
Because of the effects of fluid resistance, an object falling in a fluid does
not have a constant acceleration. To describe its motion, we can’t use the constantacceleration relationships from Chapter 2; instead, we have to start over with
Newton’s second law. As an example, suppose you drop a metal ball at the surface
of a bucket of oil and let it fall to the bottom (Fig. 5.24a). The fluid resistance
force in this situation is given by Eq. (5.5). What are the acceleration, velocity,
and position of the metal ball as functions of time?
Figure 5.24b shows the free-body diagram. We take the positive y-direction
to be downward and neglect any force associated with buoyancy in the oil. Since
the ball is moving downward, its speed v is equal to its y-velocity vy and the fluid
resistance force is in the -y@direction. There are no x-components, so Newton’s
second law gives
g Fy = mg + 1-kvy2 = may
(5.7)
BIO application Pollen and Fluid
resistance These spiky spheres are pollen grains from the ragweed flower (Ambrosia
artemisiifolia) and a common cause of hay
fever. Because of their small radius (about
10 mm = 0.01 mm), when they are released
into the air the fluid resistance force on them
is proportional to their speed. The terminal
speed given by Eq. (5.8) is only about 1 cm>s.
Hence even a moderate wind can keep pollen
grains aloft and carry them substantial distances from their source.
When the ball first starts to move, vy = 0, the resisting force is zero and the initial
acceleration is ay = g. As the speed increases, the resisting force also increases,
until finally it is equal in magnitude to the weight. At this time mg - kvy = 0,
the acceleration is zero, and there is no further increase in speed. The final speed
vt , called the terminal speed, is given by mg - kvt = 0, or
vt =
mg
k
(terminal speed, fluid resistance f = kv)
(5.8)
Figure 5.25 shows how the acceleration, velocity, and position vary with time.
As time goes by, the acceleration approaches zero and the velocity approaches vt
5.25 Graphs of the motion of a body falling without fluid resistance and with fluid resistance
proportional to the speed.
Acceleration versus time
ay
g
Velocity versus time
No fluid resistance:
vy
velocity keeps increasing.
No fluid resistance:
constant acceleration.
t
y
No fluid resistance:
parabolic curve.
vt
With fluid resistance:
position changes
more slowly.
With fluid resistance:
velocity has an upper limit.
With fluid resistance:
acceleration decreases.
O
Position versus time
O
t
O
t
5.3 Friction Forces
149
(remember that we chose the positive y-direction to be down). The slope of the
graph of y versus t becomes constant as the velocity becomes constant.
To see how the graphs in Fig. 5.25 are derived, we must find the relationship between velocity and time during the interval before the terminal speed is
reached. We go back to Newton’s second law for the falling ball, Eq. (5.7), which
we rewrite with ay = dvy>dt:
m
dvy
dt
= mg - kvy
After rearranging terms and replacing mg>k by vt , we integrate both sides, noting
that vy = 0 when t = 0:
v
dvy
L0 vy - vt
t
= -
k
dt
m L0
5.26 (a) Air drag and terminal speed.
(b) By changing the positions of their arms
and legs while falling, skydivers can
change the value of the constant D in
Eq. (5.6) and hence adjust the terminal
speed of their fall [Eq. (5.12)].
(a) Free-body diagrams for falling with air drag
which integrates to
ln
Dv2 = mg
vt - vy
vt
= -
k
t
m
or
1 -
vy
vt
Dv2 6 mg
= e-1k>m2t
ay
and finally
vy = vt 31 - e-1k>m2t4
Note that vy becomes equal to the terminal speed vt only in the limit that t S ∞;
the ball cannot attain terminal speed in any finite length of time.
The derivative of vy in Eq. (5.9) gives ay as a function of time, and the integral
of vy gives y as a function of time. We leave the derivations for you to complete;
the results are
ay = ge-1k>m2t
y = vt c t -
m
11 - e-1k>m2t2 d
k
mg
(5.9)
(5.10)
mg
y
y
Before terminal
speed: Object
accelerating, drag
force less than
weight.
At terminal speed vt :
Object in equilibrium,
drag force equals
weight.
(b) A skydiver falling at terminal speed
(5.11)
Now look again at Fig. 5.25, which shows graphs of these three relationships.
In deriving the terminal speed in Eq. (5.8), we assumed that the fluid resistance force is proportional to the speed. For an object falling through the air at
high speeds, so that the fluid resistance is equal to Dv 2 as in Eq. (5.6), the terminal speed is reached when Dv 2 equals the weight mg (Fig. 5.26a). You can show
that the terminal speed vt is given by
mg
AD
(terminal speed, fluid resistance f = Dv 2)
(5.12)
This expression for terminal speed explains why heavy objects in air tend to fall
faster than light objects. Two objects that have the same physical size but different mass (say, a table-tennis ball and a lead ball with the same radius) have the
same value of D but different values of m. The more massive object has a higher
terminal speed and falls faster. The same idea explains why a sheet of paper falls
faster if you first crumple it into a ball; the mass m is the same, but the smaller
size makes D smaller (less air drag for a given speed) and vt larger. Skydivers use
the same principle to control their descent (Fig. 5.26b).
Figure 5.27 shows the trajectories of a baseball with and without air drag,
assuming a coefficient D = 1.3 * 10-3 kg>m (appropriate for a batted ball at
sea level). Both the range of the baseball and the maximum height reached are
substantially smaller than the zero-drag calculation would lead you to believe.
Hence the baseball trajectory we calculated in Example 3.7 (Section 3.3) by
ignoring air drag is unrealistic. Air drag is an important part of the game of
baseball!
5.27 Computer-generated trajectories of a
baseball launched at 50 m>s at 35° above
the horizontal. Note that the scales are different on the horizontal and vertical axes.
50
No air drag: path is a parabola.
y (m)
vt =
With air drag: range and
maximum height are less;
path is not parabolic.
0
x (m)
250
150
Chapter 5 applying Newton’s Laws
For a human body falling through air in a spread-eagle position
(Fig. 5.26b), the numerical value of the constant D in Eq. (5.6) is
about 0.25 kg>m. Find the terminal speed for a 50-kg skydiver.
Solution
identify and Set up: This example uses the relationship among
terminal speed, mass, and drag coefficient. We use Eq. (5.12) to
find the target variable vt.
execute: We find for m = 50 kg:
vt =
Solution
tErminal SpEEd oF a SkydivEr
ExamplE 5.18
150 kg219.8 m>s22
mg
=
AD
0.25 kg>m
B
evaluate: The terminal speed is proportional to the square root
of the skydiver’s mass. A skydiver with the same drag coefficient
D but twice the mass would have a terminal speed 12 = 1.41
times greater, or 63 m>s. (A more massive skydiver would also
have more frontal area and hence a larger drag coefficient, so his
terminal speed would be a bit less than 63 m>s.) Even the 50-kg
skydiver’s terminal speed is quite high, so skydives don’t last very
long. A drop from 2800 m (9200 ft) to the surface at the terminal
speed takes only 12800 m2>144 m>s2 = 64 s.
When the skydiver deploys the parachute, the value of D
increases greatly. Hence the terminal speed of the skydiver and
parachute decreases dramatically to a much lower value.
= 44 m>s 1about 160 km>h, or 99 mi>h2
5.28 Net force, acceleration, and velocity
in uniform circular motion.
S
v
S
ΣF
S
v
S
a
S
a
In uniform
circular motion,
both acceleration
and net force are
directed toward
center of circle.
S
S
ΣF
5.4 dynamicS of circular motion
S
ΣF
a
We talked about uniform circular motion in Section 3.4. We showed that when a
particle moves in a circular path with constant speed, the particle’s acceleration
has a constant magnitude arad given by
S
v
Velocity is tangent to circle.
5.29 What happens if the inward radial
force suddenly ceases to act on a body in
circular motion?
A ball attached to a string whirls in a
circle on a frictionless surface.
S
v
S
ΣF
teSt your underStanding of Section 5.3 Consider a box that is placed
on different surfaces. (a) In which situation(s) is no friction force acting on the box?
(b) In which situation(s) is a static friction force acting on the box? (c) In which situation(s)
is a kinetic friction force acting on the box? (i) The box is at rest on a rough horizontal
surface. (ii) The box is at rest on a rough tilted surface. (iii) The box is on the roughsurfaced flat bed of a truck that is moving at a constant velocity on a straight, level road,
and the box remains in place in the middle of the truck bed. (iv) The box is on the roughsurfaced flat bed of a truck that is speeding up on a straight, level road, and the box
remains in place in the middle of the truck bed. (v) The box is on the rough-surfaced flat
bed of a truck that is climbing a hill, and the box is sliding toward the back of the truck. ❙
Suddenly, the
string breaks.
S
a
S
S
ΣF
v
S
a
S
v
S
No net force now acts on the ball, so it v
obeys Newton’s first law—it moves in a
straight line at constant velocity.
Magnitude of acceleration
of an object in
uniform circular motion
arad =
v2
R
Speed of object
Radius of object’s
circular path
(5.13)
The subscript “rad” is a reminder that at each point the acceleration points
radially inward toward the center of the circle, perpendicular to the instantaneous velocity. We explained in Section 3.4 why this acceleration is often called
centripetal acceleration or radial acceleration.
We can also express the centripetal acceleration arad in terms of the period T,
the time for one revolution:
2pR
T =
(5.14)
v
In terms of the period, arad is
Magnitude of acceleration
of an object in
uniform circular motion
arad =
4p2R
T2
Radius of object’s
circular path
(5.15)
Period of motion
Uniform circular motion, like all other motion of a particle, is governed by
Newton’s second law. STo make the particle accelerate toward the center of the
circle, the net force g F on the particle must always be directed toward the center
(Fig. 5.28). The magnitude of the acceleration is constant, so the magnitude Fnet
of the net force must also be constant. If the inward net force stops acting, the
particle flies off in a straight line tangent to the circle (Fig. 5.29).
5.4 Dynamics of Circular Motion
The magnitude of the radial acceleration is given by arad = v 2>R, so the magnitude Fnet of the net force on a particle with mass m in uniform circular motion
must be
Fnet = marad = m
5.30 Right and wrong ways to depict
uniform circular motion.
(a) Correct free-body diagram
2
v
R
(uniform circular motion)
RIGHT!
(5.16)
F
Uniform circular
motion can result from any combination of forces, just so the
S
net force g F is always directed toward the center of the circle and has a constant magnitude. Note that the body need not move around a complete circle:
Equation (5.16) is valid for any path that can be regarded as part of a circular arc.
arad
If you include the acceleration, draw it to one
side of the body to show that it’s not a force.
(b) Incorrect free-body diagram
caUtion avoid using “centrifugal force” Figure 5.30 shows a correct free-body dia-
gram for uniform circular motion (Fig. 5.30a) and an incorrect diagram (Fig. 5.30b).
Figure 5.30b is incorrect because it includes an extra outward force of magnitude m1v 2>R2
to “keep the body out there” or to “keep it in equilibrium.” There are three reasons not
to include such an outward force, called centrifugal force (“centrifugal” means “fleeing
from the center”). First, the body does not “stay out there”: It is in constant motion around
its circular path. Because its velocity is constantly changing in direction, the body accelerates and is not in equilibrium. Second, if there were an outward force that balanced the
inward force, the net force would be zero and the body would move in a straight line, not a
S
2>R2
circleS(Fig. 5.29). Third, the quantity m1v
is not a force; it corresponds to the ma side
S
S
of gF = ma and does not appear in gF (Fig. 5.30a). It’s true that when you ride in a car
that goes around a circular path, you tend to slide to the outside of the turn as though there
was a “centrifugal force.” But we saw in Section 4.2 that what happens is that you tend to
keep moving in a straight line, and the outer side of the car “runs into” you as the car turns
(Fig. 4.10c). In an inertial frame of reference there is no such thing as “centrifugal force.”
We won’t mention this term again, and we strongly advise you to avoid it. ❙
F
mv2
R
WRONG
/
The quantity mv2 R is not a force—it
doesn’t belong in a free-body diagram.
Demo
Solution
examPle 5.19
Force in uniForm circular motion
A sled with a mass of 25.0 kg rests on a horizontal sheet of essentially frictionless ice. It is attached by a 5.00-m rope to a post set
in the ice. Once given a push, the sled revolves uniformly in a
circle around the post (Fig. 5.31a). If the sled makes five complete revolutions every minute, find the force F exerted on it by
the rope.
Figure 5.31b shows our free-body diagram for the sled. The acceleration has only an x-component; this is toward the center of
the circle, so we denote it as arad. The acceleration isn’t given, so
we’ll need to determine its value by using Eq. (5.13) or Eq. (5.15).
execUte: The force F appears in Newton’s second law for the
x-direction:
gFx = F = marad
soLUtion
identiFy and set UP: The sled is in uniform circular motion, so
it has a constant radial acceleration. We’ll apply Newton’s second
law to the sled to find the magnitude F of the force exerted by the
rope (our target variable).
5.31 (a) The situation. (b) Our free-body diagram.
(a) A sled in uniform circular motion
151
(b) Free-body diagram
for sled
We can find the centripetal acceleration arad by using Eq. (5.15).
The sled moves in a circle of radius R = 5.00 m with a period
T = 160.0 s2>15 rev2 = 12.0 s, so
arad =
4p2R
T2
=
4p215.00 m2
112.0 s22
= 1.37 m>s2
The magnitude F of the force exerted by the rope is then
F = marad = 125.0 kg211.37 m>s22
= 34.3 kg # m>s2 = 34.3 N
evaLUate: You can check our value for arad by first using Eq. (5.14),
R
We point the positive
x-direction toward the
center of the circle.
v = 2pR>T, to find the speed and then using arad = v 2>R from
Eq. (5.13). Do you get the same result?
A greater force would be needed if the sled moved around
the circle at a higher speed v. In fact, if v were doubled while R
remained the same, F would be four times greater. Can you
show this? How would F change if v remained the same but the
radius R were doubled?
152
ChaPtEr 5 applying Newton’s Laws
Solution
a conical Pendulum
examPle 5.20
An inventor designs a pendulum clock using a bob with mass m at
the end of a thin wire of length L. Instead of swinging back and
forth, the bob is to move in a horizontal circle at constant speed v,
with the wire making a fixed angle b with the vertical direction
(Fig. 5.32a). This is called a conical pendulum because the suspending wire traces out a cone. Find the tension F in the wire and
the period T (the time for one revolution of the bob).
soLUtion
identiFy and set UP: To find our target variables, the tension F
and period T, we need two equations. These will be the horizontal
and vertical components of Newton’s second law applied to the
bob. We’ll find the radial acceleration of the bob from one of the
circular motion equations.
Figure 5.32b shows our free-body diagram and coordinate system for the bob at a particular instant. There are just two forces on
the bob: the weight mg and the tension F in the wire. Note that the
5.32 (a) The situation. (b) Our free-body diagram.
(a) The situation
center of the circular path is in the same horizontal plane as the
bob, not at the top end of the wire. The horizontal component of
tension is the force that produces the radial acceleration arad.
execUte: The bob has zero vertical acceleration; the horizontal
acceleration is toward the center of the circle, which is why we use
the symbol arad. Newton’s second law, Eqs. (5.2), says
gFx = F sin b = marad
gFy = F cos b + 1- mg2 = 0
These are two equations for the two unknowns F and b. The equation for gFy gives F = mg>cos b; that’s our target expression for
F in terms of b. Substituting this result into the equation for gFx
and using sin b>cos b = tan b, we find
arad = g tan b
To relate b to the period T, we use Eq. (5.15) for arad, solve for T,
and insert arad = g tan b:
arad =
4p2R
so
T2
(b) Free-body diagram
for pendulum bob
T = 2p
T2 =
4p2R
arad
R
A g tan b
Figure 5.32a shows that R = L sin b. We substitute this and use
sin b>tan b = cos b:
b
T = 2p
L
L cos b
B g
evaLUate: For a given length L, as the angle b increases,
v
R
Solution
examPle 5.21
We point the positive
x-direction toward the
center of the circle.
cos b decreases, the period T becomes smaller, and the tension
F = mg>cos b increases. The angle can never be 90°, however;
this would require that T = 0, F = q , and v = q . A conical
pendulum would not make a very good clock because the period
depends on the angle b in such a direct way.
rounding a Flat curve
The sports car in Example 3.11 (Section 3.4) is rounding a flat,
unbanked curve with radius R (Fig. 5.33a). If the coefficient of
static friction between tires and road is ms, what is the maximum
speed vmax at which the driver can take the curve without sliding?
5.33 (a) The situation. (b) Our free-body diagram.
(a) Car rounding flat curve
(b) Free-body
diagram for car
soLUtion
identiFy and set UP: The car’s acceleration as it rounds the
curve has magnitude arad = v 2>R. Hence the maximum speed vmax
(our target variable) corresponds to the maximum acceleration arad
and to the maximum horizontal force on the car toward the center of its circular path. The only horizontal force acting on the car
is the friction force exerted by the road. So to solve this problem
we’ll need Newton’s second law, the equations of uniform circular
motion, and our knowledge of the friction force from Section 5.3.
The free-body diagram in Fig. 5.33b includes the car’s weight
w = mg and the two forces exerted by the road: the normal force n
and the horizontal friction force f. The friction force must point
toward the center of the circular path in order to cause the radial
R
5.4 Dynamics of Circular Motion
153
acceleration. The car doesn’t slide toward or away from the center
of the circle, so the friction force is static friction, with a maximum magnitude fmax = msn [see Eq. (5.4)].
As an example, if ms = 0.96 and R = 230 m, we have
execute: The acceleration toward the center of the circular path
or about 170 km>h 1100 mi>h2. This is the maximum speed for
this radius.
is arad = v
2>R.
There is no vertical acceleration. Thus
gFx = f = marad = m
v2
R
gFy = n + 1- mg2 = 0
The second equation shows that n = mg. The first equation shows
that the friction force needed to keep the car moving in its circular path increases with the car’s speed. But the maximum friction
force available is fmax = ms n = ms mg, and this determines the
car’s maximum speed. Substituting ms mg for f and vmax for v in
the first equation, we find
ms mg = m
so
evaluate: If the car’s speed is slower than vmax = 2ms gR,
the required friction force is less than the maximum value
fmax = ms mg, and the car can easily make the curve. If we try to
take the curve going faster than vmax, we will skid. We could still
go in a circle without skidding at this higher speed, but the radius
would have to be larger.
The maximum centripetal acceleration (called the “lateral
acceleration” in Example 3.11) is equal to ms g. That’s why it’s
best to take curves at less than the posted speed limit if the road
is wet or icy, either of which can reduce the value of ms and
hence ms g.
vmax = 2ms gR
Solution
ExamplE 5.22
2
v max
R
vmax = 210.96219.8 m>s221230 m2 = 47 m>s
rounding a bankEd curvE
For a car traveling at a certain speed, it is possible to bank a curve
at just the right angle so that no friction is needed to maintain the
car’s turning radius. Then a car can safely round the curve even
on wet ice. (Bobsled racing depends on this idea.) Your engineering firm plans to rebuild the curve in Example 5.21 so that a car
moving at a chosen speed v can safely make the turn even with no
friction (Fig. 5.34a). At what angle b should the curve be banked?
Solution
identify and Set up: With no friction, the only forces acting
on the car are its weight and the normal force. Because the road
is banked, the normal force (which acts perpendicular to the road
surface) has a horizontal component. This component causes the
car’s horizontal acceleration toward the center of the car’s circular
path. We’ll use Newton’s second law to find the target variable b.
Our free-body diagram (Fig. 5.34b) is very similar to the diagram for the conical pendulum in Example 5.20 (Fig. 5.32b). The
normal force acting on the car plays the role of the tension force
exerted by the wire on the pendulum bob.
vertical component n cos b and a horizontal component n sin b.
The acceleration in the x-direction is the centripetal acceleration
arad = v 2>R; there is no acceleration in the y-direction. Thus the
equations of Newton’s second law are
gFx = n sin b = marad
gFy = n cos b + 1- mg2 = 0
From the gFy equation, n = mg>cos b. Substituting this into the
gFx equation and using arad = v 2>R, we get an expression for the
bank angle:
arad
v2
v2
tan b =
so
b = arctan
=
g
gR
gR
evaluate: The bank angle depends on both the speed and the radius. For a given radius, no one angle is correct for all speeds. In the
design of highways and railroads, curves are often banked for the average speed of the traffic over them. If R = 230 m and v = 25 m>s
(equal to a highway speed of 88 km>h, or 55 mi>h), then
b = arctan
S
execute: The normal force n is perpendicular to the roadway
and is at an angle b with the vertical (Fig. 5.34b). Thus it has a
125 m>s22
19.8 m>s221230 m2
This is within the range of bank angles actually used in highways.
5.34 (a) The situation. (b) Our free-body diagram.
(a) Car rounding banked curve
Vertical
b
Normal to road is at
same angle from the
vertical as road is
from the horizontal.
b
Horizontal
R
= 15°
(b) Free-body diagram for car
154
Chapter 5 applying Newton’s Laws
5.35 An airplane banks to one side in
order to turn in that direction.
The vertical
S
component of the lift force L balances the
force
of gravity; the horizontal component
S
of L causes the acceleration v 2>R.
L cos b
L
b
L sin b
w = mg
Banked curves and the flight of airplanes
The results of Example 5.22 also apply to an airplane when it makes a turn in
level flight (Fig. 5.35). When an airplane is flying in a straight line at a constant
speed Sand at a steady altitude, the airplane’s weight is exactly balanced by the lift
force L exerted by the air. (The upward lift force that the air exerts on the wings
is a reaction to the downward push the wings exert on the air as they move
through it.) To make the airplane turn, the pilot banks the airplane to one side
so that the lift force has a horizontal component, as Fig. 5.35 shows. (The pilot
also changes the angle at which the wings “bite” into the air so that the vertical
component of lift continues to balance the weight.) The bank angle is related to
the airplane’s speed v and the radius R of the turn by the same expression as in
Example 5.22: tan b = v 2>gR. For an airplane to make a tight turn (small R) at
high speed (large v), tan b must be large and the required bank angle b must
approach 90°.
We can also apply the results of Example 5.22 to the pilot of an airplane. The
free-body diagram for the pilot of the airplane is exactly as shown in Fig. 5.34b; the
normal force n = mg>cos b is exerted on the pilot by the seat. As in Example 5.9,
n is equal to the apparent weight of the pilot, which is greater than the pilot’s true
weight mg. In a tight turn with a large bank angle b, the pilot’s apparent weight can
be tremendous: n = 5.8mg at b = 80° and n = 9.6mg at b = 84°. Pilots black
out in such tight turns because the apparent weight of their blood increases by the
same factor, and the human heart isn’t strong enough to pump such apparently
“heavy” blood to the brain.
motion in a vertical circle
In Examples 5.19, 5.20, 5.21, and 5.22 the body moved in a horizontal circle.
Motion in a vertical circle is no different in principle, but the weight of the body
has to be treated carefully. The following example shows what we mean.
Solution
ExamplE 5.23
uniForm circular motion in a vErtical circlE
A passenger on a carnival Ferris wheel moves in a vertical circle
of radius R with constant speed v. The seat remains upright during
the motion. Find expressions for the force the seat exerts on the
passenger when at the top of the circle and when at the bottom.
5.36 Our sketches for this problem.
(a) Sketch of two positions
Solution
(b) Free-body diagram
for passenger at top
(c) Free-body diagram
for passenger at bottom
identify and Set up: The target variables are nT, the upward
normal force the seat applies to the passenger at the top of the
circle, and nB, the normal force at the bottom. We’ll find these
by using Newton’s second law and the uniform circular motion
equations.
Figure 5.36a shows the passenger’s velocity and acceleration
at the two positions. The acceleration always points toward the
center of the circle—downward at the top of the circle and upward
at the bottom of the circle. At each position the only forces acting
are vertical: the upward normal force and the downward force of
gravity. Hence we need only the vertical component of Newton’s
second law. Figures 5.36b and 5.36c show free-body diagrams for
the two positions. We take the positive y-direction as upward in
both cases (that is, opposite the direction of the acceleration at the
top of the circle).
execute: At the top the acceleration has magnitude v 2>R, but its
vertical component is negative because its direction is downward.
Hence ay = - v 2>R and Newton’s second law tells us that
Top:
a Fy = nT + 1-mg2 = - m
nT = mg a1 -
v2
b
gR
v2
R
or
5.5 the Fundamental Forces of Nature
At the bottom the acceleration is upward, so ay = + v 2>R and
Newton’s second law says
Bottom:
gFy = nB + 1- mg2 = +m
nB = mg a1 +
v2
R
or
v2
b
gR
evaluate: Our result for nT tells us that at the top of the Ferris
wheel, the upward force the seat applies to the passenger is smaller
155
in magnitude than the passenger’s weight w = mg. If the ride goes
fast enough that g - v 2>R becomes zero, the seat applies no force,
and the passenger is about to become airborne. If v becomes still
larger, nT becomes negative; this means that a downward force
(such as from a seat belt) is needed to keep the passenger in the
seat. By contrast, the normal force nB at the bottom is always
greater than the passenger’s weight. You feel the seat pushing up
on you more firmly than when you are at rest. You can see that nT
and nB are the values of the passenger’s apparent weight at the top
and bottom of the circle (see Section 5.2).
When we tie a string to an object and whirl it in a vertical circle, the analysis
in Example 5.23 isn’t directly applicable. The reason is that v is not constant in
this case; except at the top and bottom of the circle, the net force (and hence the
acceleration)
does not point toward the center of the circle (Fig. 5.37). So both
S
S
g F and a have a component tangent to the circle, which means that the speed
changes. Hence this is a case of nonuniform circular motion (see Section 3.4).
Even worse, we can’t use the constant-acceleration formulas to relate the speeds
at various points because neither the magnitude nor the direction of the acceleration is constant. The speed relationships we need are best obtained by using the
concept of energy. We’ll consider such problems in Chapter 7.
teSt your underStanding of Section 5.4 Satellites are held in orbit by
the force of our planet’s gravitational attraction. A satellite in a small-radius orbit moves
at a higher speed than a satellite in an orbit of large radius. Based on this information,
what can you conclude about the earth’s gravitational attraction for the satellite? (i) It
increases with increasing distance from the earth. (ii) It is the same at all distances from
the earth. (iii) It decreases with increasing distance from the earth. (iv) This information
by itself isn’t enough to answer the question. ❙
5.5 the fundamental forceS of nature
We have discussed several kinds of forces—including weight, tension, friction,
fluid resistance, and the normal force—and we will encounter others as we
continue our study of physics. How many kinds of forces are there? Our best
understanding is that all forces are expressions of just four distinct classes of
fundamental forces, or interactions between particles (Fig. 5.38, next page).
Two are familiar in everyday experience. The other two involve interactions between subatomic particles that we cannot observe with the unaided senses.
Gravitational interactions include the familiar force of your weight, which
results from the earth’s gravitational attraction acting on you. The mutual gravitational attraction of various parts of the earth for each other holds our planet
together, and likewise for the other planets (Fig. 5.38a). Newton recognized that
the sun’s gravitational attraction for the earth keeps our planet in its nearly circular orbit around the sun. In Chapter 13 we’ll study gravitational interactions in
more detail, including their vital role in the motions of planets and satellites.
The second familiar class of forces, electromagnetic interactions, includes
electric and magnetic forces. If you run a comb through your hair, the comb ends
up with an electric charge; you can use the electric force exerted by this charge
to pick up bits of paper. All atoms contain positive and negative electric charge,
so atoms and molecules can exert electric forces on one another. Contact forces,
including the normal force, friction, and fluid resistance, are the result of electrical interactions between atoms on the surface of an object and atoms in its
surroundings (Fig. 5.38b). Magnetic forces, such as those between magnets or
between a magnet and a piece of iron, are actually the result of electric charges
in motion. For example, an electromagnet causes magnetic interactions because
5.37 A ball moving in a vertical circle.
When a ball moves in a vertical circle ...
... the net force on the ball has
a component toward the center
of the circle ...
T
... but also a component
tangent to the circle ...
a ... so the net acceleration
is not purely radial.
w = mg
BIO application circular motion in a
centrifuge An important tool in medicine
and biological research is the ultracentrifuge,
a device that makes use of the dynamics of
circular motion. A tube is filled with a solvent
that contains various small particles (for example, blood containing platelets and white
and red blood cells). The tube is inserted into
the centrifuge, which then spins at thousands
of revolutions per minute. The solvent provides the inward force that keeps the particles
in circular motion. The particles slowly drift
away from the rotation axis within the solvent.
Because the drift rate depends on the particle
size and density, particles of different types
become separated in the tube, making analysis much easier.
156
ChaPtEr 5 applying Newton’s Laws
5.38 Examples of the fundamental
interactions in nature.
(a) The gravitational interaction
Saturn is held together by the mutual
gravitional attraction of all of its parts.
(b) The electromagnetic interaction
The contact forces between the microphone
and the singer’s hand are electrical in nature.
The particles that make up the rings are
held in orbit by Saturn’s gravitational force.
This microphone uses electric and magnetic
effects to convert sound into an electrical
signal that can be amplified and recorded.
(c) The strong interaction
The nucleus of a gold atom has 79 protons
and 118 neutrons.
(d) The weak interaction
Scientists find the age of this ancient
skull by measuring its carbon-14—a
form of carbon that is radioactive thanks
to the weak interaction.
The strong interaction holds the protons
and neutrons together and overcomes the
electric repulsion of the protons.
electric charges move through its wires. We will study electromagnetic interactions in detail in the second half of this book.
On the atomic or molecular scale, gravitational forces play no role because
electric forces are enormously stronger: The electrical repulsion between two
protons is stronger than their gravitational attraction by a factor of about 1035.
But in bodies of astronomical size, positive and negative charges are usually
present in nearly equal amounts, and the resulting electrical interactions nearly
cancel out. Gravitational interactions are thus the dominant influence in the
motion of planets and in the internal structure of stars.
The other two classes of interactions are less familiar. One, the strong interaction, is responsible for holding the nucleus of an atom together (Fig. 5.38c).
Nuclei contain electrically neutral neutrons and positively charged protons. The
electric force between charged protons tries to push them apart; the strong attractive force between nuclear particles counteracts this repulsion and makes the nucleus stable. In this context the strong interaction is also called the strong nuclear
force. It has much shorter range than electrical interactions, but within its range
it is much stronger. Without the strong interaction, the nuclei of atoms essential
to life, such as carbon (six protons, six neutrons) and oxygen (eight protons, eight
neutrons), would not exist and you would not be reading these words!
Finally, there is the weak interaction. Its range is so short that it plays a role
only on the scale of the nucleus or smaller. The weak interaction is responsible
for a common form of radioactivity called beta decay, in which a neutron in a
radioactive nucleus is transformed into a proton while ejecting an electron and
a nearly massless particle called an antineutrino. The weak interaction between
157
Summary
the antineutrino and ordinary matter is so feeble that an antineutrino could easily
penetrate a wall of lead a million kilometers thick!
An important application of the weak interaction is radiocarbon dating, a
technique that enables scientists to determine the ages of many biological specimens (Fig. 5.38d). Naturally occurring carbon includes atoms of both carbon-12
(with six protons and six neutrons in the nucleus) and carbon-14 (with two
additional neutrons). Living organisms take in carbon atoms of both kinds from
their environment but stop doing so when they die. The weak interaction makes
carbon-14 nuclei unstable—one of the neutrons changes to a proton, an electron,
and an antineutrino—and these nuclei decay at a known rate. By measuring the
fraction of carbon-14 that is left in an organism’s remains, scientists can determine how long ago the organism died.
In the 1960s physicists developed a theory that described the electromagnetic
and weak interactions as aspects of a single electroweak interaction. This theory
has passed every experimental test to which it has been put. Encouraged by this
success, physicists have made similar attempts to describe the strong, electromagnetic, and weak interactions in terms of a single grand unified theory (GUT)
and have taken steps toward a possible unification of all interactions into a theory
of everything (TOE). Such theories are still speculative, and there are many unanswered questions in this very active field of current research.
Chapter
5 Summary
Using Newton’s first law: When a body is in equilibrium in an inertial frame of reference—that is,
either at rest or moving with constant velocity—the
vector sum of forces acting on it must be zero
(Newton’s first law). Free-body diagrams are
essential in identifying the forces that act on the
body being considered.
Newton’s third law (action and reaction) is also
frequently needed in equilibrium problems. The
two forces in an action–reaction pair never act on
the same body. (See Examples 5.1–5.5.)
The normal force exerted on a body by a
surface is not always equal to the body’s weight.
(See Example 5.4.)
Using Newton’s second law: If the vector sum of
forces on a body is not zero, the body accelerates.
The acceleration is related to the net force by
Newton’s second law.
Just as for equilibrium problems, free-body diagrams are essential for solving problems involving
Newton’s second law, and the normal force exerted
on a body is not always equal to its weight. (See
Examples 5.6–5.12.)
SolutionS to all exampleS
y
Vector form:
gF = 0
S
Component form:
gFx = 0
gFy = 0
n
n
(5.1)
T
w sin a
T
w cos a
a
x
a
w
w
Vector form:
y
gF = ma
S
S
Component form:
gFx = max
gFy = may
a
(5.2)
n
n
ax
T
m
a
w
w sin a
T
w cos a
x
a
w
158
Chapter 5 applying Newton’s Laws
Magnitude of kinetic friction force:
between two bodies can always be represented
fk = mk n
(5.3)
S
in terms of a normal force n perpendicular
to the
S
Magnitude of static friction force:
surface of contact and a friction force f parallel to
the surface.
fs … 1fs2max = msn
(5.4)
When a body is sliding over the surface, the friction force is called kinetic friction. Its magnitude fk is
approximately equal to the normal force magnitude n
multiplied by the coefficient of kinetic friction mk .
When a body is not moving relative to a surface, the friction force is called static friction. The
maximum possible static friction force is approximately equal to the magnitude n of the normal force
multiplied by the coefficient of static friction ms . The actual static friction force may be anything from
zero to this maximum value, depending on the situation. Usually ms is greater than mk for a given pair
of surfaces in contact. (See Examples 5.13–5.17.)
Rolling friction is similar to kinetic friction, but the force of fluid resistance depends on the
speed of an object through a fluid. (See Example 5.18.)
Friction and fluid resistance: The contact force
Forces in circular motion: In uniform circular
Acceleration in uniform circular motion:
motion, the acceleration vector is directed toward
the center of the circle. The Smotion is governed
S
by Newton’s second law, gF = ma. (See
Examples 5.19–5.23.)
arad =
Solution guide
identify and Set up
1. Although we want the block not to slide up or down on the
inside of the cone, this is not an equilibrium problem. The block
rotates with the cone and is in uniform circular motion, so it has
an acceleration directed toward the center of its circular path.
2. Identify the forces on the block. What is the direction of the friction force when the cone is rotating as slowly as possible, so T
5.39 A block inside a spinning cone.
R
fk
T
O
S
v
ΣF
S
v
S
arad
S
arad
S
ΣF
S
ΣF
in a rotating conE
A small block with mass m is placed inside an inverted cone that is
rotating about a vertical axis such that the time for one revolution
of the cone is T (Fig. 5.39). The walls of the cone make an angle b
with the horizontal. The coefficient of static friction between the
block and the cone is ms. If the block is to remain at a constant
height h above the apex of the cone, what are (a) the maximum
value of T and (b) the minimum value of T? (That is, find expressions for Tmax and Tmin in terms of b and h.)
1 fs2max
S
(5.13), (5.15)
Kinetic
friction
S
arad
S
v
Solution
bridging problEm
v2
4p2R
=
R
T2
Static
friction
f
has its maximum value Tmax? What is the direction of the friction force when the cone is rotating as rapidly as possible, so T
has its minimum value Tmin? In these situations does the static
friction force have its maximum magnitude? Why or why not?
3. Draw a free-body diagram for the block when the cone is rotating with T = Tmax and a free-body diagram when the cone
is rotating with T = Tmin. Choose coordinate axes, and remember that it’s usually easiest to choose one of the axes to
be in the direction of the acceleration.
4. What is the radius of the circular path that the block follows?
Express this in terms of b and h.
5. List the unknown quantities, and decide which of these are
the target variables.
execute
6. Write Newton’s second law in component form for the case in
which the cone is rotating with T = Tmax. Write the acceleration in terms of Tmax, b, and h, and write the static friction
force in terms of the normal force n.
7. Solve these equations for the target variable Tmax.
8. Repeat steps 6 and 7 for the case in which the cone is rotating
with T = Tmin, and solve for the target variable Tmin.
evaluate
m
h
b
Time for 1 rotation = T
9. You’ll end up with some fairly complicated expressions for
Tmax and Tmin, so check them over carefully. Do they have the
correct units? Is the minimum time Tmin less than the maximum time Tmax, as it must be?
10. What do your expressions for Tmax and Tmin become if ms = 0?
Check your results by comparing them with Example 5.22 in
Section 5.4.
Discussion Questions
159
Problems
For assigned homework and other learning materials, go to MasteringPhysics®.
., .., ...: Difficulty levels. CP: Cumulative problems incorporating material from earlier chapters. CALC: Problems requiring calculus.
DATA: Problems involving real data, scientific evidence, experimental design, and/or statistical reasoning. BIO: Biosciences problems.
[Always assume that pulleys are frictionless and massless and that strings and cords are massless, unless otherwise noted.]
discUssion qUestions
Q5.1 A man sits in a seat that is hanging from a rope. The rope
passes over a pulley suspended from the ceiling, and the man
holds the other end of the rope in his hands. What is the tension in
the rope, and what force does the seat exert on him? Draw a freebody force diagram for the man.
Q5.2 “In general, the normal force is not equal to the weight.”
Give an example in which these two forces are equal in magnitude, and at least two examples in which they are not.
Q5.3 A clothesline hangs between two poles. No matter how
tightly the line is stretched, it sags a little at the center. Explain why.
Q5.4 You drive a car up a steep hill at constant speed. Discuss all
of the forces that act on the car. What pushes it up the hill?
Q5.5 For medical reasons, astronauts in outer space must determine their body mass at regular intervals. Devise a scheme for
measuring body mass in an apparently weightless environment.
Q5.6 To push a box up a ramp, which requires less force: pushing
horizontally or pushing parallel to the ramp? Why?
Q5.7 A woman in an elevator lets go of her briefcase, but it does
not fall to the floor. How is the elevator moving?
Q5.8 A block rests on an inclined plane with enough friction to
prevent it from sliding down. To start the block moving, is it easier
to push it up the plane or down the plane? Why?
Q5.9 A crate slides up an inclined ramp and then slides down the
ramp after momentarily stopping near the top. There is kinetic
friction between the surface of the ramp and the crate. Which is
greater? (i) The crate’s acceleration going up the ramp; (ii) the crate’s
acceleration going down the ramp; (iii) both are the same. Explain.
Q5.10 A crate of books rests on a level floor. To move it along
the floor at a constant velocity, why do you exert less force if you
pull it at an angle u above the horizontal than if you push it at the
same angle below the horizontal?
Q5.11 In a world without friction, which of the following activities could you do (or not do)? Explain your reasoning. (a) Drive
around an unbanked highway curve; (b) jump into the air; (c) start
walking on a horizontal sidewalk; (d) climb a vertical ladder;
(e) change lanes while you drive.
Q5.12 When you stand with bare feet in a wet bathtub, the grip
feels fairly secure, and yet a catastrophic slip is quite possible.
Explain this in terms of the two coefficients of friction.
Q5.13 You are pushing a large crate from the back of a freight
elevator to the front as the elevator is moving to the next floor. In
which situation is the force you must apply to move the crate the
least, and in which is it the greatest: when the elevator is accelerating upward, when it is accelerating downward, or when it is traveling at constant speed? Explain.
Q5.14 It is often said that “friction always opposes motion.”
Give at least one example in which (a) static friction causes motion,
and (b) kinetic friction causes motion.
Q5.15 If there is a net force on a particle in uniform circular motion, why doesn’t the particle’s speed change?
Q5.16 A curve in a road has a bank angle calculated and posted
for 80 km>h. However, the road is covered with ice, so you cautiously plan to drive slower than this limit. What might happen to
your car? Why?
Q5.17 You swing a ball on the end of a lightweight string in a
horizontal circle at constant speed. Can the string ever be truly
horizontal? If not, would it slope above the horizontal or below
the horizontal? Why?
Q5.18 The centrifugal force is not included in the free-body diagrams of Figs. 5.34b and 5.35. Explain why not.
Q5.19 A professor swings a rubber stopper in a horizontal circle
on the end of a string in front of his class. He tells Caroline, in the
front row, that he is going to let the string go when the stopper is
directly in front of her face. Should Caroline worry?
Q5.20 To keep the forces on the riders within allowable limits,
many loop-the-loop roller coaster rides are designed so that the
loop is not a perfect circle but instead has a larger radius of curvature at the bottom than at the top. Explain.
Q5.21 A tennis ball drops from rest at the top of a tall glass
cylinder—first with the air pumped out of the cylinder so that
there is no air resistance, and again after the air has been readmitted to the cylinder. You examine multiflash photographs of
the two drops. Can you tell which photo belongs to which drop?
If so, how?
Q5.22 You throw a baseball straight upward with speed v0.
When the ball returns to the point from where you threw it, how
does its speed compare to v0 (a) in the absence of air resistance
and (b) in the presence of air resistance? Explain.
Q5.23 You throw a baseball straight upward. If you do not ignore
air resistance, how does the time required for the ball to reach
its maximum height compare to the time required for it to fall
from its maximum height back down to the height from which you
threw it? Explain.
Q5.24 You have two identical tennis balls and fill one with
water. You release both balls simultaneously from the top of a tall
building. If air resistance is negligible, which ball will strike the
ground first? Explain. What if air resistance is not negligible?
Q5.25 A ball is dropped from rest and feels air resistance as it
falls. Which of the graphs in Fig. Q5.25 best represents its acceleration as a function of time?
Figure Q5.25
a
a
a
a
t
t
t
t
(b)
(a)
a
(c)
t
(e)
(d)
Q5.26 A ball is dropped from rest and feels air resistance as it
falls. Which of the graphs in Fig. Q5.26 best represents its vertical velocity component as a function of time?
Figure Q5.26
v
v
t
(a)
v
v
t
(b)
v
t
t
(c)
(d)
t
(e)
160
ChaPtEr 5 applying Newton’s Laws
Q5.27 When a batted baseball moves with air drag, when does
the ball travel a greater horizontal distance? (i) While climbing
to its maximum height; (ii) while descending from its maximum
height back to the ground; (iii) the same for both? Explain in terms
of the forces acting on the ball.
Q5.28 “A ball is thrown from the edge of a high cliff. Regardless
of the angle at which it is thrown, due to air resistance, the ball
will eventually end up moving vertically downward.” Justify this
statement.
exercises
Section 5.1 Using Newton’s First law: Particles
in equilibrium
5.1 . Two 25.0-N weights are suspended at opposite ends of a rope
that passes over a light, frictionless pulley. The pulley is attached
to a chain from the ceiling. (a) What is the tension in the rope?
(b) What is the tension in the chain?
5.2 . In Fig. E5.2 each of the suspended blocks has weight w.
The pulleys are frictionless, and the ropes have negligible weight.
In each case, draw a free-body diagram and calculate the tension T
in the rope in terms of w.
.. A picture frame hung against a wall is suspended by two
wires attached to its upper corners. If the two wires make the
same angle with the vertical, what must this angle be if the tension
in each wire is equal to 0.75 of the weight of the frame? (Ignore
any friction between the wall and the picture frame.)
5.6 .. A large wrecking ball Figure e5.6
is held in place by two light
steel cables (Fig. E5.6). If the
mass m of the wrecking ball is
3620 kg, what are (a) the tension TB in the cable that makes
40°
TB
an angle of 40° with the vertical and (b) the tension TA in the
horizontal cable?
TA
m
5.5
.. Find the tension in each cord in Fig. E5.7 if the weight of
the suspended object is w.
5.7
Figure e5.7
(a)
Figure e5.2
(a)
(b)
30° 45°
B
A
45°
B
60°
C
(b)
C
A
(c)
w
w
.. A 1130-kg car is held Figure e5.8
in place by a light cable on a
very smooth (frictionless) ramp
(Fig. E5.8). The cable makes
an angle of 31.0° above the sur31.0°
face of the ramp, and the ramp
itself rises at 25.0° above the
horizontal. (a) Draw a free-body
diagram for the car. (b) Find
25.0°
the tension in the cable. (c) How
hard does the surface of the
ramp push on the car?
5.9 .. A man pushes on a piano with mass 180 kg; it slides at
constant velocity down a ramp that is inclined at 19.0° above the
horizontal floor. Neglect any friction acting on the piano.
Calculate the magnitude of the force applied by the man if he
pushes (a) parallel to the incline and (b) parallel to the floor.
5.10 .. In Fig. E5.10 the weight w is 60.0 N. (a) What is the tension in the diagonal
string?
(b) Find the magnitudes of the horiS
S
zontal forces F1 and F2 that must be applied to hold the system in
the position shown.
5.8
w
ble
w
w
w
5.3 . A 75.0-kg wrecking ball hangs from a uniform, heavy-duty
chain of mass 26.0 kg. (a) Find the maximum and minimum tensions in the chain. (b) What is the tension at a point three-fourths
of the way up from the bottom of the chain?
5.4 .. BIO Injuries to the Spinal Column. In the treatment of
spine injuries, it is often necessary to provide tension along the
spinal column to stretch the backbone. One device for doing this is
the Stryker frame (Fig. E5.4a). A weight W is attached to the
patient (sometimes around a neck collar, Fig. E5.4b), and friction
between the person’s body and the bed prevents sliding. (a) If the
coefficient of static friction between a 78.5-kg patient’s body and
the bed is 0.75, what is the maximum traction force along the
spinal column that W can provide without causing the patient to
slide? (b) Under the conditions of maximum traction, what is the
tension in each cable attached to the neck collar?
Ca
w
Figure e5.10
Figure e5.4
(a)
(b)
90.0°
S
F1
W
90.0°
45.0°
S
90.0°
65°
65°
w
F2
exercises
Section 5.2 Using Newton’s Second Law:
Dynamics of Particles
..
BIO Stay Awake! An astronaut is inside a 2.25 * 106 kg
rocket that is blasting off vertically from the launch pad. You want
this rocket to reach the speed of sound 1331 m>s2 as quickly as
possible, but astronauts are in danger of blacking out at an acceleration greater than 4g. (a) What is the maximum initial thrust this
rocket’s engines can have but just barely avoid blackout? Start
with a free-body diagram of the rocket. (b) What force, in terms of
the astronaut’s weight w, does the rocket exert on her? Start with a
free-body diagram of the astronaut. (c) What is the shortest time it
can take the rocket to reach the speed of sound?
5.12 .. A rocket of initial mass 125 kg (including all the contents) has an engine that produces a constant vertical force (the
thrust) of 1720 N. Inside this rocket, a 15.5-N electrical power
supply rests on the floor. (a) Find the initial acceleration of the
rocket. (b) When the rocket initially accelerates, how hard does
the floor push on the power supply? (Hint: Start with a free-body
diagram for the power supply.)
5.13 .. CP Genesis Crash. On September 8, 2004, the Genesis
spacecraft crashed in the Utah desert because its parachute did not
open. The 210-kg capsule hit the ground at 311 km>h and penetrated the soil to a depth of 81.0 cm. (a) What was its acceleration
(in m>s2 and in g’s), assumed to be constant, during the crash?
(b) What force did the ground exert on the capsule during the
crash? Express the force in newtons and as a multiple of the capsule’s weight. (c) How long did this force last?
5.14 . Three sleds are being pulled horizontally on frictionless
horizontal ice using horizontal ropes (Fig. E5.14). The pull is of
magnitude 190 N. Find (a) the acceleration of the system and (b) the
tension in ropes A and B.
5.11
Figure E5.14
30.0 kg
B
20.0 kg
A
10.0 kg
Pull
Figure E5.15
5.15 .. Atwood’s Machine. A
15.0-kg load of bricks hangs
from one end of a rope that
passes over a small, frictionless
pulley. A 28.0-kg counterweight
is suspended from the other end
of the rope (Fig. E5.15). The
system is released from rest.
(a) Draw two free-body diagrams, one for the load of bricks
28.0 kg
and one for the counterweight.
(b) What is the magnitude of the
upward acceleration of the load
of bricks? (c) What is the tension
in the rope while the load is
15.0 kg
moving? How does the tension
compare to the weight of the
load of bricks? To the weight of the counterweight?
5.16 .. CP An 8.00-kg block of ice, released from rest at the top
of a 1.50-m-long frictionless ramp, slides downhill, reaching a
speed of 2.50 m>s at the bottom. (a) What is the angle between the
ramp and the horizontal? (b) What would be the speed of the ice at
the bottom if the motion were opposed by a constant friction force
of 10.0 N parallel to the surface of the ramp?
161
5.17 .. A light rope is attached to a block with mass 4.00 kg that
rests on a frictionless, horizontal surface. The horizontal rope
passes over a frictionless, massless pulley, and a block with mass m
is suspended from the other end. When the blocks are released,
the tension in the rope is 15.0 N. (a) Draw two free-body diagrams:
one for each block. (b) What is the acceleration of either block?
(c) Find m. (d) How does the tension compare to the weight of the
hanging block?
5.18 .. CP Runway Design. A transport plane takes off from
a level landing field with two gliders in tow, one behind the other.
The mass of each glider is 700 kg, and the total resistance (air
drag plus friction with the runway) on each may be assumed constant and equal to 2500 N. The tension in the towrope between
the transport plane and the first glider is not to exceed 12,000 N.
(a) If a speed of 40 m>s is required for takeoff, what minimum
length of runway is needed? (b) What is the tension in the towrope between the two gliders while they are accelerating for the
takeoff?
5.19 .. CP A 750.0-kg boulder is raised from a quarry 125 m
deep by a long uniform chain having a mass of 575 kg. This chain
is of uniform strength, but at any point it can support a maximum
tension no greater than 2.50 times its weight without breaking.
(a) What is the maximum acceleration the boulder can have and
still get out of the quarry, and (b) how long does it take to be lifted
out at maximum acceleration if it started from rest?
5.20 .. Apparent Weight. A 550-N physics student stands on
a bathroom scale in an elevator that is supported by a cable. The
combined mass of student plus elevator is 850 kg. As the elevator
starts moving, the scale reads 450 N. (a) Find the acceleration of
the elevator (magnitude and direction). (b) What is the acceleration if the scale reads 670 N? (c) If the scale reads zero, should the
student worry? Explain. (d) What is the tension in the cable in
parts (a) and (c)?
5.21 .. CP BIO Force During a Jump. When jumping straight
up from a crouched position, an average person can reach a maximum height of about 60 cm. During the jump, the person’s body
from the knees up typically rises a distance of around 50 cm.
To keep the calculations simple and yet get a reasonable result,
assume that the entire body rises this much during the jump.
(a) With what initial speed does the person leave the ground to
reach a height of 60 cm? (b) Draw a free-body diagram of the person during the jump. (c) In terms of this jumper’s weight w, what
force does the ground exert on him or her during the jump?
5.22 CP CALC A 2540-kg test rocket is launched vertically from
the launch pad. Its fuel (of negligible mass) provides a thrust force
such that its vertical velocity as a function of time is given by
v1t2 = At + Bt 2, where A and B are constants and time is measured from the instant the fuel is ignited. The rocket has an
upward acceleration of 1.50 m>s2 at the instant of ignition and,
1.00 s later, an upward velocity of 2.00 m>s. (a) Determine A and B,
including their SI units. (b) At 4.00 s after fuel ignition, what is
the acceleration of the rocket, and (c) what thrust force does the
burning fuel exert on it, assuming no air resistance? Express the
thrust in newtons and as a multiple of the rocket’s weight. (d) What
was the initial thrust due to the fuel?
5.23 .. CP CALC A 2.00-kg box is moving to the right with
speed 9.00 m>s on a horizontal, frictionless surface. At t = 0 a
horizontal force is applied to the box. The force is directed to the
left and has magnitude F1t2 = 16.00 N>s22t 2. (a) What distance
does the box move from its position at t = 0 before its speed is
reduced to zero? (b) If the force continues to be applied, what is
the speed of the box at t = 3.00 s?
162
5.24
Chapter 5 applying Newton’s Laws
..
CP CALC A 5.00-kg crate is suspended from the end of
a short vertical rope of negligible mass. An upward force F1t2 is
applied to the end of the rope, and the height of the crate above its
initial position is given by y1t2 = 12.80 m>s2t + 10.610 m>s32t 3.
What is the magnitude of F when t = 4.00 s?
Section 5.3 Friction Forces
5.25 . BIO The Trendelenburg Position. After emergencies
with major blood loss, a patient is placed in the Trendelenburg position, in which the foot of the bed is raised to get maximum blood
flow to the brain. If the coefficient of static friction between a
typical patient and the bedsheets is 1.20, what is the maximum
angle at which the bed can be tilted with respect to the floor before the patient begins to slide?
5.26 . In a laboratory experiment on friction, a 135-N block
resting on a rough horizontal table is pulled by a horizontal wire.
The pull gradually increases until the block begins to move and
continues to increase thereafter. Figure E5.26 shows a graph of
the friction force on this block as a function of the pull. (a) Identify
the regions of the graph where static friction and kinetic friction
occur. (b) Find the coefficients of static friction and kinetic friction between the block and the table. (c) Why does the graph slant
upward at first but then level out? (d) What would the graph look
like if a 135-N brick were placed on the block, and what would the
coefficients of friction be?
Figure E5.26
f (N)
75.0
50.0
25.0
O
5.27
25.0 50.0 75.0 100.0 125.0 150.0
P (N)
to give it an acceleration of 1.10 m>s2? (c) Suppose you were performing the same experiment on the moon, where the acceleration
due to gravity is 1.62 m>s2. (i) What magnitude push would cause
it to move? (ii) What would its acceleration be if you maintained
the push in part (b)?
5.30 .. Some sliding rocks approach the base of a hill with a
speed of 12 m>s. The hill rises at 36° above the horizontal and has
coefficients of kinetic friction and static friction of 0.45 and 0.65,
respectively, with these rocks. (a) Find the acceleration of the
rocks as they slide up the hill. (b) Once a rock reaches its highest
point, will it stay there or slide down the hill? If it stays, show
why. If it slides, find its acceleration on the way down.
5.31 .. A box with mass 10.0 kg moves on a ramp that is inclined at an angle of 55.0o above the horizontal. The coefficient of
kinetic friction between the box and the ramp surface is
mk = 0.300. Calculate the magnitude of the acceleration of the
box if you push on the box with a constant force F = 120.0 N that
is parallel to the ramp surface and (a) directed down the ramp,
moving the box down the ramp; (b) directed up the ramp, moving
the box up the ramp.
5.32 .. A pickup truck is carrying a toolbox, but the rear gate of
the truck is missing. The toolbox will slide out if it is set moving.
The coefficients of kinetic friction and static friction between the
box and the level bed of the truck are 0.355 and 0.650, respectively.
Starting from rest, what is the shortest time this truck could
accelerate uniformly to 30.0 m>s without causing the box to slide?
Draw a free-body diagram of the toolbox.
5.33 .. You are lowering two boxes, one on top of the other,
down a ramp by pulling on a rope parallel to the surface of the
ramp (Fig. E5.33). Both boxes move together at a constant speed
of 15.0 cm>s. The coefficient of kinetic friction between the ramp
and the lower box is 0.444, and the coefficient of static friction
between the two boxes is 0.800. (a) What force do you need to
exert to accomplish this? (b) What are the magnitude and direction of the friction force on the upper box?
Figure E5.33
0
32.
kg
..
CP A stockroom worker pushes a box with mass 16.8 kg
on a horizontal surface with a constant speed of 3.50 m>s. The
coefficient of kinetic friction between the box and the surface is
0.20. (a) What horizontal force must the worker apply to maintain
the motion? (b) If the force calculated in part (a) is removed, how
far does the box slide before coming to rest?
5.28 .. A box of bananas weighing 40.0 N rests on a horizontal
surface. The coefficient of static friction between the box and the
surface is 0.40, and the coefficient of kinetic friction is 0.20. (a) If
no horizontal force is applied to the box and the box is at rest, how
large is the friction force exerted on it? (b) What is the magnitude
of the friction force if a monkey applies a horizontal force of 6.0 N
to the box and the box is initially at rest? (c) What minimum horizontal force must the monkey apply to start the box in motion?
(d) What minimum horizontal force must the monkey apply to
keep the box moving at constant velocity once it has been started?
(e) If the monkey applies a horizontal force of 18.0 N, what is the
magnitude of the friction force and what is the box’s acceleration?
5.29 .. A 45.0-kg crate of tools rests on a horizontal floor. You
exert a gradually increasing horizontal push on it, and the crate
just begins to move when your force exceeds 313 N. Then you
must reduce your push to 208 N to keep it moving at a steady
25.0 cm>s. (a) What are the coefficients of static and kinetic friction between the crate and the floor? (b) What push must you exert
0
48.
kg
2.50 m
4.75 m
5.34
..
Consider the system shown in Fig. E5.34. Block A weighs
45.0 N, and block B weighs 25.0 N. Once block B is set into downward motion, it descends at a constant speed. (a) Calculate the
coefficient of kinetic friction between block A and the tabletop.
(b) A cat, also of weight 45.0 N, falls asleep on top of block A. If
block B is now set into downward motion, what is its acceleration
(magnitude and direction)?
Figure E5.34
A
B
exercises
5.35 .. CP Stopping Distance. (a) If the coefficient of kinetic
friction between tires and dry pavement is 0.80, what is the shortest distance in which you can stop a car by locking the brakes
when the car is traveling at 28.7 m>s (about 65 mi>h)? (b) On wet
pavement the coefficient of kinetic friction may be only 0.25. How
fast should you drive on wet pavement to be able to stop in the
same distance as in part (a)? (Note: Locking the brakes is not the
safest way to stop.)
5.36 .. CP A 25.0-kg box of textbooks rests on a loading ramp
that makes an angle a with the horizontal. The coefficient of kinetic friction is 0.25, and the coefficient of static friction is 0.35.
(a) As a is increased, find the minimum angle at which the box
starts to slip. (b) At this angle, find the acceleration once the box
has begun to move. (c) At this angle, how fast will the box be moving after it has slid 5.0 m along the loading ramp?
5.37 . Two crates connected by a rope lie on a horizontal surface (Fig. E5.37). Crate A has mass m A, and crate B has mass m B.
The coefficient of kinetic friction between each crate and the surface is mk. The crates
are pulled to the right at constant velocity by
S
a horizontal force F. Draw one or more free-body diagrams to
calculateS the following in terms of m A, m B, and mk: (a) the magnitude of F and (b) the tension in the rope connecting the blocks.
S
A
5.38
Section 5.4 Dynamics of Circular Motion
. A stone with mass 0.80 kg is attached to one end of a
string 0.90 m long. The string will break if its tension exceeds
60.0 N. The stone is whirled in a horizontal circle on a frictionless
tabletop; the other end of the string remains fixed. (a) Draw a freebody diagram of the stone. (b) Find the maximum speed the stone
can attain without the string breaking.
5.44 . BIO Force on a Skater’s Wrist. A 52-kg ice skater spins
about a vertical axis through her body with her arms horizontally
outstretched; she makes 2.0 turns each second. The distance from
one hand to the other is 1.50 m. Biometric measurements indicate
that each hand typically makes up about 1.25% of body weight.
(a) Draw a free-body diagram of one of the skater’s hands. (b) What
horizontal force must her wrist exert on her hand? (c) Express the
force in part (b) as a multiple of the weight of her hand.
5.45 .. A small remote-controlled car with mass 1.60 kg moves
at a constant speed of v = 12.0 m>s in a track formed by a vertical
circle inside a hollow metal cylinder that has a radius of 5.00 m
(Fig. E5.45). What is the magnitude of the normal force exerted
on the car by the walls of the cylinder at (a) point A (bottom of the
track) and (b) point B (top of the track)?
5.43
Figure E5.45
Figure E5.37
B
163
B
F
5.00 m
..
A box with mass m is dragged across a level floor with
coefficient of kinetic friction mk by a rope that is pulled upward
at an angle u above the horizontal with a force of magnitude F.
(a) In terms of m, mk, u, and g, obtain an expression for the magnitude of the force required to move the box with constant speed.
(b) Knowing that you are studying physics, a CPR instructor asks
you how much force it would take to slide a 90-kg patient across a
floor at constant speed by pulling on him at an angle of 25° above
the horizontal. By dragging weights wrapped in an old pair of
pants down the hall with a spring balance, you find that mk = 0.35.
Use the result of part (a) to answer the instructor’s question.
5.39 .. CP As shown in Fig. E5.34, block A (mass 2.25 kg) rests
on a tabletop. It is connected by a horizontal cord passing over a
light, frictionless pulley to a hanging block B (mass 1.30 kg). The
coefficient of kinetic friction between block A and the tabletop is
0.450. The blocks are released then from rest. Draw one or more
free-body diagrams to find (a) the speed of each block after they
move 3.00 cm and (b) the tension in the cord.
5.40 .. You throw a baseball straight upward. The drag force is
proportional to v 2. In terms of g, what is the y-component of the
ball’s acceleration when the ball’s speed is half its terminal speed
and (a) it is moving up? (b) It is moving back down?
5.41 .. A large crate with mass m rests on a horizontal floor. The
coefficients of friction between the crate and
the floor are ms and mk.
S
A woman pushes downward with a force F on the crate
at an angle u
S
below the horizontal. (a) What magnitude of force F is required to
keep the crate moving at constant velocity? (b) If ms is greater than
some critical value, the woman cannot start the crate moving no
matter how hard she pushes. Calculate this critical value of ms.
5.42 . (a) In Example 5.18 (Section 5.3), what value of D is
required to make vt = 42 m>s for the skydiver? (b) If the skydiver’s
daughter, whose mass is 45 kg, is falling through the air and has
the same D 10.25 kg>m2 as her father, what is the daughter’s terminal speed?
A
5.46 .. A small car with mass 0.800 kg travels at constant speed
on the inside of a track that is a vertical circle with radius 5.00 m
(Fig. E5.45). If the normal force exerted by the track on the car
when it is at the top of the track (point B) is 6.00 N, what is the normal force on the car when it is at the bottom of the track (point A)?
5.47 . A small model car with mass m travels at constant speed
on the inside of a track that is a vertical circle with radius 5.00 m
(Fig. E5.45). If the normal force exerted by the track on the car
when it is at the bottom of the track (point A) is equal to 2.50mg,
how much time does it take the car to complete one revolution
around the track?
5.48 . A flat (unbanked) curve on a highway has a radius of
170.0 m. A car rounds the curve at a speed of 25.0 m>s. (a) What
is the minimum coefficient of static friction that will prevent
sliding? (b) Suppose that the highway is icy and the coefficient of
static friction between the tires and pavement is only one-third of
what you found in part (a). What should be the maximum speed
of the car so that it can round the curve safely?
5.49 .. A 1125-kg car and a 2250-kg pickup truck approach a
curve on a highway that has a radius of 225 m. (a) At what angle
should the highway engineer bank this curve so that vehicles traveling at 65.0 mi>h can safely round it regardless of the condition of
their tires? Should the heavy truck go slower than the lighter car?
(b) As the car and truck round the curve at 65.0 mi>h, find the
normal force on each one due to the highway surface.
164
ChaPtEr 5 applying Newton’s Laws
5.50 .. The “Giant Swing” at a county fair consists of a vertical
central shaft with a number of horizontal arms attached at its
upper end. Each arm supports a seat suspended from a cable 5.00 m
long, and the upper end of the cable is fastened to the arm at a
point 3.00 m from the central shaft (Fig. E5.50). (a) Find the time
of one revolution of the swing if the cable supporting a seat makes
an angle of 30.0° with the vertical. (b) Does the angle depend on
the weight of the passenger for a given rate of revolution?
Figure e5.50
3.00 m
5.0
0m
30.0°
5.51 .. In another version of the Figure e5.51
“Giant Swing” (see Exercise 5.50),
the seat is connected to two cables, one of which is horizontal
40.0°
(Fig. E5.51). The seat swings in
a horizontal circle at a rate of
28.0 rpm 1rev>min2. If the seat
weighs 255 N and an 825-N person is sitting in it, find the ten7.50 m
sion in each cable.
5.52
..
A small button placed on a horizontal rotating platform
with diameter 0.520 m will revolve with the platform when it is
brought up to a speed of 40.0 rev>min, provided the button is no
more than 0.220 m from the axis. (a) What is the coefficient of
static friction between the button and the platform? (b) How far
from the axis can the button be placed, without slipping, if the
platform rotates at 60.0 rev>min?
5.53 .. Rotating Space Stations. One problem for humans
living in outer space is that they are apparently weightless. One
way around this problem is to design a space station that spins
about its center at a constant rate. This creates “artificial gravity”
at the outside rim of the station. (a) If the diameter of the space
station is 800 m, how many revolutions per minute are needed for
the “artificial gravity” acceleration to be 9.80 m>s2? (b) If the
space station is a waiting area for travelers going to Mars, it might
be desirable to simulate the acceleration due to gravity on the
Martian surface 13.70 m>s22. How many revolutions per minute
are needed in this case?
5.54 . The Cosmo Clock 21 Ferris wheel in Yokohama, Japan,
has a diameter of 100 m. Its name comes from its 60 arms, each of
which can function as a second hand (so that it makes one revolution every 60.0 s). (a) Find the speed of the passengers when the
Ferris wheel is rotating at this rate. (b) A passenger weighs 882 N
at the weight-guessing booth on the ground. What is his apparent
weight at the highest and at the lowest point on the Ferris wheel?
(c) What would be the time for one revolution if the passenger’s
apparent weight at the highest point were zero? (d) What then
would be the passenger’s apparent weight at the lowest point?
5.55 .. An airplane flies in a loop (a circular path in a vertical
plane) of radius 150 m. The pilot’s head always points toward the
center of the loop. The speed of the airplane is not constant;
the airplane goes slowest at the top of the loop and fastest at the
bottom. (a) What is the speed of the airplane at the top of the loop,
where the pilot feels weightless? (b) What is the apparent weight
of the pilot at the bottom of the loop, where the speed of the airplane is 280 km>h? His true weight is 700 N.
5.56 .. A 50.0-kg stunt pilot who has been diving her airplane
vertically pulls out of the dive by changing her course to a circle
in a vertical plane. (a) If the plane’s speed at the lowest point of the
circle is 95.0 m>s, what is the minimum radius of the circle so that
the acceleration at this point will not exceed 4.00g? (b) What is
the apparent weight of the pilot at the lowest point of the pullout?
5.57 . Stay Dry! You tie a cord to a pail of water and swing
the pail in a vertical circle of radius 0.600 m. What minimum
speed must you give the pail at the highest point of the circle to
avoid spilling water?
5.58 .. A bowling ball weighing 71.2 N 116.0 lb2 is attached to
the ceiling by a 3.80-m rope. The ball is pulled to one side and
released; it then swings back and forth as a pendulum. As the
rope swings through the vertical, the speed of the bowling ball
is 4.20 m>s. At this instant, what are (a) the acceleration of the
bowling ball, in magnitude and direction, and (b) the tension in
the rope?
5.59 .. BIO Effect on Blood of Walking. While a person is
walking, his arms swing through approximately a 45° angle in 12 s.
As a reasonable approximation, assume that the arm moves with
constant speed during each swing. A typical arm is 70.0 cm long,
measured from the shoulder joint. (a) What is the acceleration of a
1.0-g drop of blood in the fingertips at the bottom of the swing?
(b) Draw a free-body diagram of the drop of blood in part (a).
(c) Find the force that the blood vessel must exert on the drop of
blood in part (a). Which way does this force point? (d) What force
would the blood vessel exert if the arm were not swinging?
ProbLems
5.60 .. An adventurous archaeologist crosses between two rock
cliffs by slowly going hand over hand along a rope stretched
between the cliffs. He stops to rest at the middle of the rope
(Fig. P5.60). The rope will break if the tension in it exceeds
2.50 * 104 N, and our hero’s mass is 90.0 kg. (a) If the angle u
is 10.0°, what is the tension in the rope? (b) What is the smallest
value u can have if the rope is not to break?
Figure P5.60
u
u
165
Problems
Figure P5.63
B
A
D
C
37.1°
... A horizontal wire holds a solid uniform ball of mass m
in place on a tilted ramp that rises 35.0° above the horizontal. The
surface of this ramp is perfectly smooth, and the wire is directed
away from the center of the ball (Fig. P5.64). (a) Draw a freebody diagram of the ball. (b) How hard does the surface of the
ramp push on the ball? (c) What is the tension in the wire?
5.64
Figure P5.64
35.0°
Figure P5.65
cm
... A solid uniform 45.0-kg ball of
diameter 32.0 cm is supported against a
vertical, frictionless wall by a thin 30.0-cm
wire of negligible mass (Fig. P5.65). (a) Draw
a free-body diagram for the ball, and use
the diagram to find the tension in the wire.
(b) How hard does the ball push against the
wall?
5.65
30.0
... Two ropes are connected Figure P5.61
to a steel cable that supports a hanging weight (Fig. P5.61). (a) Draw a
60° 40°
free-body diagram showing all of
the forces acting at the knot that
connects the two ropes to the steel
cable. Based on your diagram,
which of the two ropes will have the
greater tension? (b) If the maximum
tension either rope can sustain without breaking is 5000 N, determine the maximum value of the hanging weight that these ropes
can safely support. Ignore the weight of the ropes and of the steel
cable.
5.62 .. In Fig. P5.62 a worker lifts a Figure P5.62
weight w by pulling
down on a rope
S
with a force F. The upper pulley is
attached to the ceiling by a chain, and
the lower pulley is attached to the
weight by another chain. Draw one or
more free-body diagrams to find the
tension in
each chain and the magniS
S
tude of F, in terms of w, if the weight
F
is lifted at constant speed. Assume
that the rope, pulleys, and chains have
w
negligible weights.
5.63 .. In a repair shop a truck engine that has mass 409 kg is
held in place by four light cables (Fig. P5.63). Cable A is horizontal, cables B and D are vertical, and cable C makes an angle of
37.1o with a vertical wall. If the tension in cable A is 722 N, what
are the tensions in cables B and C?
5.61
5.66 .. CP A box is sliding with a constant speed of 4.00 m>s in
the + x-direction on a horizontal, frictionless surface. At x = 0
the box encounters a rough patch of the surface, and then the surface becomes even rougher. Between x = 0 and x = 2.00 m, the
coefficient of kinetic friction between the box and the surface is
0.200; between x = 2.00 m and x = 4.00 m, it is 0.400. (a) What
is the x-coordinate of the point where the box comes to rest?
(b) How much time does it take the box to come to rest after it
first encounters the rough patch at x = 0?
5.67 .. CP BIO Forces During Chin-ups. When you do a
chin-up, you raise your chin just over a bar (the chinning bar), supporting yourself with only your arms. Typically, the body below
the arms is raised by about 30 cm in a time of 1.0 s, starting from
rest. Assume that the entire body of a 680-N person doing chinups is raised by 30 cm, and that half the 1.0 s is spent accelerating
upward and the other half accelerating downward, uniformly in
both cases. Draw a free-body diagram of the person’s body, and
use it to find the force his arms must exert on him during the
accelerating part of the chin-up.
5.68 .. CP CALC A 2.00-kg box is suspended from the end of a
light vertical rope. A time-dependent force is applied to the upper
end of the rope, and the box moves upward with a velocity magnitude that varies in time according to v1t2 = 12.00 m>s22t +
10.600 m>s32t 2. What is the tension in the rope when the velocity
of the box is 9.00 m>s?
5.69 ... CALC A 3.00-kg box that is several hundred meters
above the earth’s surface is suspended from the end of a short
vertical rope of negligible mass. A time-dependent upward force is
applied to the upper end of the rope and results in a tension in the
rope of T1t2 = 136.0 N>s2t. The box is at rest at t = 0. The only
forces on the box are the tension in the rope and gravity. (a) What
is the velocity of the box at (i) t = 1.00 s and (ii) t = 3.00 s?
(b) What is the maximum distance that the box descends below
its initial position? (c) At what value of t does the box return to its
initial position?
5.70 .. CP A 5.00-kg box sits at rest at the bottom of a ramp that
is 8.00 m long and is inclined at 30.0o above the horizontal. The
coefficient of kinetic friction is mk = 0.40, and the coefficient of
static friction is ms = 0.43. What constant force F, applied parallel to the surface of the ramp, is required to push the box to the top
of the ramp in a time of 6.00 s?
5.71 .. Two boxes connected by a light horizontal rope are on a
horizontal surface (Fig. E5.37). The coefficient of kinetic friction between each box and the surface is mk = 0.30. Box B has
mass 5.00 kg, and box A has mass m. A force F with magnitude
40.0 N and direction 53.1° above the horizontal is applied to the
5.00-kg box, and both boxes move to the right with a = 1.50 m>s2.
(a) What is the tension T in the rope that connects the boxes?
(b) What is m?
166
Chapter 5 applying Newton’s Laws
5.72 ... A 6.00-kg box sits on a ramp that is inclined at 37.0°
above the horizontal. The coefficient of kinetic friction between the
box and the ramp is mk = 0.30. What horizontal force is required
to move the box up the incline with a constant acceleration of
3.60 m>s2?
5.73 .. CP An 8.00-kg box sits Figure P5.73
on a ramp that is inclined at 33.0°
above the horizontal. The coeffiF
cient of kinetic friction between the
box and the surface of the ramp is
mk = 0.300. A constant horizontal
force F = 26.0 N is applied to the
33.0°
box (Fig. P5.73), and the box
moves down the ramp. If the box is
initially at rest, what is its speed
2.00 s after the force is applied?
5.74 .. CP In Fig. P5.74, m 1 = 20.0 kg and a = 53.1°. The coefficient of kinetic friction between the block of mass m 1 and the
incline is mk = 0.40. What must be the mass m 2 of the hanging
block if it is to descend 12.0 m in the first 3.00 s after the system is
released from rest?
.. A block with mass m 1 is placed on an inclined plane with
slope angle a and is connected to a hanging block with mass m 2
by a cord passing over a small, frictionless pulley (Fig. P5.74). The
coefficient of static friction is ms , and the coefficient of kinetic
friction is mk. (a) Find the value of m 2 for which the block of
mass m 1 moves up the plane at constant speed once it is set in motion. (b) Find the value of m 2 for which the block of mass m 1 moves
down the plane at constant speed once it is set in motion. (c) For
what range of values of m 2 will the blocks remain at rest if they
are released from rest?
5.78 .. BIO The Flying Leap of a Flea. High-speed motion
pictures 13500 frames>second2 of a jumping 210@mg flea yielded
the data to plot the flea’s acceleration as a function of time, as
shown in Fig. P5.78. (See “The Flying Leap of the Flea,” by
M. Rothschild et al., Scientific American, November 1973.) This
flea was about 2 mm long and jumped at a nearly vertical takeoff
angle. Using the graph, (a) find the initial net external force on
the flea. How does it compare to the flea’s weight? (b) Find the
maximum net external force on this jumping flea. When does
this maximum force occur? (c) Use the graph to find the flea’s
maximum speed.
Figure P5.74
Figure P5.78
5.77
150
m1
100
a>g
m2
a
50
5.75 .. CP You place a book of mass 5.00 kg
againstS a vertical wall. You apply a constant
force F to the book, where F = 96.0 N and the
force is at an angle of 60.0o above the horizontal
(Fig. P5.75). The coefficient of kinetic friction
between the book and the wall is 0.300. If the
book is initially at rest, what is its speed after it
has traveled 0.400 m up the wall?
Figure P5.75
0
0
0.5
1.0
1.5
Time (ms)
S
F
5.79 .. Block A in Fig. P5.79 weighs 1.20 N, and block B weighs
3.60 N. The coefficient of kinetic friction betweenSall surfaces is
0.300. Find the magnitude of the horizontal force F necessary to
drag block B to the left at constant speed (a) if A rests on B and
moves with it (Fig. P5.79a), (b) if A is held at rest (Fig. P5.79b).
60.0°
5.76
Figure P5.79
..
Block A in Fig. P5.76 weighs 60.0 N. The coefficient of
static friction between the block and the surface on which it rests
is 0.25. The weight w is 12.0 N, and the system is in equilibrium.
(a) Find the friction force exerted on block A. (b) Find the maximum weight w for which the system will remain in equilibrium.
Figure P5.76
A
45.0°
w
(a)
(b)
A
F
A
S
S
B
F
B
5.80 ... CP Elevator Design. You are designing an elevator
for a hospital. The force exerted on a passenger by the floor of the
elevator is not to exceed 1.60 times the passenger’s weight. The
elevator accelerates upward with constant acceleration for a distance of 3.0 m and then starts to slow down. What is the maximum speed of the elevator?
5.81 ... CP CALC You are standing on a bathroom scale in an
elevator in a tall building. Your mass is 64 kg. The elevator starts
from rest and travels upward with a speed that varies with time
according to v1t2 = 13.0 m>s22t + 10.20 m>s32t 2. When t = 4.0 s,
what is the reading on the bathroom scale?
Problems
5.82 .. A hammer is hanging by a light rope from the ceiling of
a bus. The ceiling is parallel to the roadway. The bus is traveling
in a straight line on a horizontal street. You observe that the hammer hangs at rest with respect to the bus when the angle between
the rope and the ceiling of the bus is 56.0°. What is the acceleration of the bus?
5.83 .. A 40.0-kg packing case is initially at rest on the floor of
a 1500-kg pickup truck. The coefficient of static friction between
the case and the truck floor is 0.30, and the coefficient of kinetic
friction is 0.20. Before each acceleration given below, the truck is
traveling due north at constant speed. Find the magnitude and direction of the friction force acting on the case (a) when the truck
accelerates at 2.20 m>s2 northward and (b) when it accelerates at
3.40 m>s2 southward.
5.84 ... If the coefficient of static friction between a table and a
uniform, massive rope is ms, what fraction of the rope can hang
over the edge of the table without the rope sliding?
5.85 ... Two identical 15.0-kg balls,
Figure P5.85
each 25.0 cm in diameter, are suspended
by two 35.0-cm wires (Fig. P5.85). The
entire apparatus is supported by a
18.0 cm
single 18.0-cm wire, and the surfaces
of the balls are perfectly smooth.
(a) Find the tension in each of the three
wires. (b) How hard does each ball
35.0 cm
35.0 cm
push on the other one?
167
if block B is moving to the right and speeding up with an acceleration of 2.00 m>s2? (b) What is the tension in each cord when block B
has this acceleration?
Figure P5.89
B
S
a
C
A
5.90 .. Two blocks connected by a cord passing over a small,
frictionless pulley rest on frictionless planes (Fig. P5.90). (a) Which
way will the system move when the blocks are released from rest?
(b) What is the acceleration of the blocks? (c) What is the tension
in the cord?
Figure P5.90
100 kg
50 kg
30.0°
5.86
.
CP Traffic Court. You are called as an expert witness
in a trial for a traffic violation. The facts are these: A driver
slammed on his brakes and came to a stop with constant acceleration. Measurements of his tires and the skid marks on the pavement indicate that he locked his car’s wheels, the car traveled 192 ft
before stopping, and the coefficient of kinetic friction between the
road and his tires was 0.750. He was charged with speeding in a
45@mi>h zone but pleads innocent. What is your conclusion: guilty
or innocent? How fast was he going when he hit his brakes?
5.87 ... Block A in Fig. P5.87
Figure P5.87
weighs 1.90 N, and block B
A
weighs 4.20 N. The coefficient
of kinetic friction between all
S
surfaces is 0.30. Find the magB
F
S
nitude of the horizontal force F
necessary to drag block B to the
left at constant speed if A and B
are connected by a light, flexible cord passing around a fixed,
frictionless pulley.
5.88 .. CP Losing Cargo. A 12.0-kg box rests on the level
bed of a truck. The coefficients of friction between the box and
bed are ms = 0.19 and mk = 0.15. The truck stops at a stop sign
and then starts to move with an acceleration of 2.20 m>s2. If the
box is 1.80 m from the rear of the truck when the truck starts, how
much time elapses before the box falls off the truck? How far does
the truck travel in this time?
5.89 .. Block A in Fig. P5.89 has mass 4.00 kg, and block B has
mass 12.0 kg. The coefficient of kinetic friction between block B
and the horizontal surface is 0.25. (a) What is the mass of block C
53.1°
5.91 .. In terms of m 1, m 2, and g, find the acceleration of each
block in Fig. P5.91. There is no friction anywhere in the system.
Figure P5.91
m1
m2
5.92 ... Block B, with mass Figure P5.92
5.00 kg, rests on block A, with
B
mass 8.00 kg, which in turn
A
is on a horizontal tabletop
(Fig. P5.92). There is no friction between block A and the
tabletop, but the coefficient of
C
static friction between blocks
A and B is 0.750. A light string
attached to block A passes over
a frictionless, massless pulley, and block C is suspended from the
other end of the string. What is the largest mass that block C can
have so that blocks A and B still slide together when the system
is released from rest?
168
Chapter 5 applying Newton’s Laws
... Two objects, with masses 5.00 kg and 2.00 kg, hang
0.600 m above the floor from the ends of a cord that is 6.00 m
long and passes over a frictionless pulley. Both objects start from
rest. Find the maximum height reached by the 2.00-kg object.
5.94 .. Friction in an Elevator. You are riding in an elevator
on the way to the 18th floor of your dormitory. The elevator is
accelerating upward with a = 1.90 m>s2. Beside you is the box
containing your new computer; the box and its contents have a
total mass of 36.0 kg. While the elevator is accelerating upward,
you push horizontally on the box to slide it at constant speed
toward the elevator door. If the coefficient of kinetic friction
between the box and the elevator floor is mk = 0.32, what magnitude of force must you apply?
5.95 . A block is placed Figure P5.95
S
against the vertical front of a
a
cart (Fig. P5.95). What acceleration must the cart have so
that block A does not fall? The
coefficient of static friction beA
tween the block and the cart is
ms. How would an observer on
the cart describe the behavior
of the block?
5.96 ... Two blocks, with masses 4.00 kg and 8.00 kg, are connected by a string and slide down a 30.0° inclined plane
(Fig. P5.96). The coefficient of kinetic friction between the 4.00-kg
block and the plane is 0.25; that between the 8.00-kg block and
the plane is 0.35. Calculate (a) the acceleration of each block and
(b) the tension in the string. (c) What happens if the positions of
the blocks are reversed, so that the 4.00-kg block is uphill from
the 8.00-kg block?
5.99 ... Banked Curve I. A curve with a 120-m radius on a
level road is banked at the correct angle for a speed of 20 m>s. If
an automobile rounds this curve at 30 m>s, what is the minimum
coefficient of static friction needed between tires and road to
prevent skidding?
5.100 .. Banked Curve II. Consider a wet roadway banked
as in Example 5.22 (Section 5.4), where there is a coefficient of
static friction of 0.30 and a coefficient of kinetic friction of 0.25
between the tires and the roadway. The radius of the curve is
R = 50 m. (a) If the bank angle is b = 25°, what is the maximum
speed the automobile can have before sliding up the banking?
(b) What is the minimum speed the automobile can have before
sliding down the banking?
5.101 ... Blocks A, B, and C are placed as in Fig. P5.101 and
connected by ropes of negligible mass. Both A and B weigh
25.0 N each, and the coefficient of kinetic friction between each
block and the surface is 0.35. Block C descends with constant
velocity. (a) Draw separate free-body diagrams showing the forces
acting on A and on B. (b) Find the tension in the rope connecting
blocks A and B. (c) What is the weight of block C? (d) If the
rope connecting A and B were cut, what would be the acceleration of C?
Figure P5.96
.. You are riding in a school bus. As the bus rounds a flat
curve at constant speed, a lunch box with mass 0.500 kg, suspended from theUniversity
ceiling of
the bus
Physics
14e by a string 1.80 m long, is
found to hang at Young/Freedman
rest relative to the bus when the string makes an
Benjamin
Cummings
angle of 30.0° with
the vertical.
In this position the lunch box is
50.0 m from the Pearson
curve’s Education
center of curvature. What is the speed v
9736105103
of the bus?
Fig 05_P101
..
5.103
CALC Pickup:
You throw
a rock downward into water with a
6754605084
Rolin Graphics
speed of 3mg>k, where
k is the coefficient in Eq. (5.5). Assume that
jr 3/26/14fluid
16p0
x 7p5
the relationship between
resistance
and speed is as given
in Eq. (5.5), and calculate the speed of the rock as a function
of time.
5.104 ... A 4.00-kg block is Figure P5.104
attached to a vertical rod by
means of two strings. When the
system rotates about the axis of
the rod, the strings are extended
1.25 m
as shown in Fig. P5.104 and
the tension in the upper string
2.00 m
4.00 kg
is 80.0 N. (a) What is the tension in the lower cord? (b) How
1.25 m
many revolutions per minute
does the system make? (c) Find
the number of revolutions per
minute at which the lower cord
just goes slack. (d) Explain what
happens if the number of revolutions per minute is less than
that in part (c).
5.93
8.00
kg
4.00
kg
30°
5.97
...
Block A, with weight 3w, slides down an inclined plane S
of slope angle 36.9° at a constant Figure P5.97
speed while plank B, with
weight w, rests on top of A.
The plank is attached by a
cord to the wall (Fig. P5.97).
B
A
(a) Draw a diagram of all the
forces acting on block A. (b) If
the coefficient of kinetic friction is the same between A and B
and between S and A, deter36.9°
S
mine its value.
5.98 .. Jack sits in the chair
of a Ferris wheel that is rotating at a constant 0.100 rev>s. As Jack
passes through the highest point of his circular path, the upward
force that the chair exerts on him is equal to one-fourth of his
weight. What is the radius of the circle in which Jack travels?
Treat him as a point mass.
Figure P5.101
B
A
5.102
C
36.9°
problems
.. On the ride “Spindletop” at the amusement park Six
Flags Over Texas, people stood against the inner wall of a hollow
vertical cylinder with radius 2.5 m. The cylinder started to rotate,
and when it reached a constant rotation rate of 0.60 rev>s, the
floor dropped about 0.5 m. The people remained pinned against
the wall without touching the floor. (a) Draw a force diagram for a
person on this ride after the floor has dropped. (b) What minimum coefficient of static friction was required for the person not
to slide downward to the new position of the floor? (c) Does your
answer in part (b) depend on the person’s mass? (Note: When such
a ride is over, the cylinder is slowly brought to rest. As it slows
down, people slide down the walls to the floor.)
5.106 .. A 70-kg person rides in a 30-kg cart moving at 12 m>s
at the top of a hill that is in the shape of an arc of a circle with a
radius of 40 m. (a) What is the apparent weight of the person as
the cart passes over the top of the hill? (b) Determine the maximum speed that the cart can travel at the top of the hill without
losing contact with the surface. Does your answer depend on the
mass of the cart or the mass of the person? Explain.
5.107 .. A small bead can
slide without friction on a circu- Figure P5.107
lar hoop that is in a vertical plane
and has a radius of 0.100 m.
The hoop rotates at a constant
rate of 4.00 rev>s about a vertical diameter (Fig. P5.107).
(a) Find the angle b at which
the bead is in vertical equilib0.100 m
rium. (It has a radial acceleration toward the axis.) (b) Is it
possible for the bead to “ride”
b
at the same elevation as the
center of the hoop? (c) What will
happen if the hoop rotates at
1.00 rev>s?
5.105
.. A physics major is working to pay her college tuition by
performing in a traveling carnival. She rides a motorcycle inside a
hollow, transparent plastic sphere. After gaining sufficient speed,
she travels in a vertical circle with radius 13.0 m. She has mass
70.0 kg, and her motorcycle has mass 40.0 kg. (a) What minimum
speed must she have at the top of the circle for the motorcycle tires
to remain in contact with the sphere? (b) At the bottom of the
circle, her speed is twice the value calculated in part (a). What is
the magnitude of the normal force exerted on the motorcycle by
the sphere at this point?
5.109 .. DATA In your physics lab, a block of mass m is at rest
on a horizontal surface. You attach a light cord to the block and
apply a horizontal force to the free end of the cord. You find that
the block remains at rest until the tension T in the cord exceeds
20.0 N. For T 7 20.0 N, you measure the acceleration of the block
when T is maintained at a constant value, and you plot the results
(Fig. P5.109). The equation for the straight line that best fits your
data is a = 30.182 m>1N # s224T - 2.842 m>s2. For this block
and surface, what are (a) the coefficient of static friction and
(b) the coefficient of kinetic friction? (c) If the experiment were
done on the earth’s moon, where g is much smaller than on the
earth, would the graph of a versus T still be fit well by a straight
line? If so, how would the slope and intercept of the line differ
from the values in Fig. P5.109? Or, would each of them be the
same?
5.108
169
Figure P5.109
a (m>s2)
7.00
6.00
5.00
4.00
3.00
2.00
1.00
T (N)
0.00
20.0 25.0 30.0 35.0 40.0 45.0 50.0 55.0
.. DATA A road heading due east passes over a small hill.
You drive a car of mass m at constant speed v over the top of the
hill, where the shape of the roadway is well approximated as an
arc of a circle with radius R. Sensors have been placed on the road
surface there to measure the downward force that cars exert on
the surface at various speeds. The table gives values of this force
versus speed for your car:
5.110
Speed 1m , s 2
Force 1N2
6.00
8.00
10.0
12.0
14.0
16.0
8100
7690
7050
6100
5200
4200
Treat the car as a particle. (a) Plot the values in such a way that
they are well fitted by a straight line. You might need to raise the
speed, the force, or both to some power. (b) Use your graph from
part (a) to calculate m and R. (c) What maximum speed can the car
have at the top of the hill and still not lose contact with the road?
5.111 .. DATA You are an engineer working for a manufacturing company. You are designing a mechanism that uses a cable to
drag heavy metal blocks a distance of 8.00 m along a ramp that is
sloped at 40.0° above the horizontal. The coefficient of kinetic
friction between these blocks and the incline is mk = 0.350. Each
block has a mass of 2170 kg. The block will be placed on the bottom of the ramp, the cable will be attached, and the block will
then be given just enough of a momentary push to overcome static
friction. The block is then to accelerate at a constant rate to move
the 8.00 m in 4.20 s. The cable is made of wire rope and is parallel
to the ramp surface. The table gives the breaking strength of the
cable as a function of its diameter; the safe load tension, which is
20% of the breaking strength; and the mass per meter of the cable:
Cable Diameter Breaking Strength
(kN)
(in.)
Safe Load Mass per Meter
(kN)
(kg/m)
1
4
24.4
4.89
0.16
3
8
54.3
10.9
0.36
1
2
95.2
19.0
0.63
5
8
149
29.7
0.98
3
4
212
42.3
1.41
7
8
286
57.4
1.92
1
372
74.3
2.50
Source: www.engineeringtoolbox.com
170
ChaPtEr 5 applying Newton’s Laws
(a) What is the minimum diameter of the cable that can be used to
pull a block up the ramp without exceeding the safe load value of
the tension in the cable? Ignore the mass of the cable, and select
the diameter from those listed in the table. (b) You need to know
safe load values for diameters that aren’t in the table, so you hypothesize that the breaking strength and safe load limit are proportional to the cross-sectional area of the cable. Draw a graph
that tests this hypothesis, and discuss its accuracy. What is your
9
estimate of the safe load value for a cable with diameter 16
in.?
(c) The coefficient of static friction between the crate and the ramp
is ms = 0.620, which is nearly twice the value of the coefficient
of kinetic friction. If the machinery jams and the block stops in
the middle of the ramp, what is the tension in the cable? Is it
larger or smaller than the value when the block is moving? (d) Is
the actual tension in the cable, at its upper end, larger or smaller
than the value calculated when you ignore the mass of the cable?
If the cable is 9.00 m long, how accurate is it to ignore the cable’s
mass?
5.115 ... A ball is held at rest at position A in Fig. P5.115 by
two light strings. The horizontal string is cut, and the ball starts
swinging as a pendulum. Position B is the farthest to the right that
the ball can go as it swings back and forth. What is the ratio of the
tension in the supporting string at B to its value at A before the
string was cut?
Figure P5.115
b
A
b
B
chaLLenge ProbLems
5.112 ... Moving Wedge. A wedge with mass M rests on a
frictionless, horizontal tabletop. A block with mass m is placed on
the wedge (Fig. P5.112a). There is no friction between the block
and the wedge. The system is released from rest. (a) Calculate the
acceleration of the wedge and the horizontal and vertical components of the acceleration of the block. (b) Do your answers to part (a)
reduce to the correct results when M is very large? (c) As seen by a
stationary observer, what is the shape of the trajectory of the block?
Figure P5.112
m
(a)
m
(b)
S
F
M
M
a
a
5.113 ... A wedge with mass M rests on a frictionless, horizontal tabletop. A block
with mass m is placed on the wedge, and a
S
horizontal force F is applied
to the wedge (Fig. P5.112b). What
S
must the magnitude of F be if the block is to remain at a constant
height above the tabletop?
5.114 ... Double Atwood’s Figure P5.114
Machine. In Fig. P5.114
masses m 1 and m 2 are connected by a light string A over
a light, frictionless pulley B.
The axle of pulley B is connected by a light string C over
D
C
a light, frictionless pulley D to
a mass m 3 . Pulley D is suspended from the ceiling by an
attachment to its axle. The
B
m3
system is released from rest. In
A
terms of m 1, m 2, m 3, and g,
m2
what are (a) the acceleration of
m
1
block m 3; (b) the acceleration
of pulley B; (c) the acceleration of block m 1; (d) the acceleration of block m 2; (e) the tension in string A; (f) the tension in
string C? (g) What do your expressions give for the special case of
m 1 = m 2 and m 3 = m 1 + m 2? Is this reasonable?
Passage ProbLems
FricTion and climbing ShoES. Shoes made for the sports
of bouldering and rock climbing are designed to provide a great
deal of friction between the foot and the surface of the ground.
Such shoes on smooth rock might have a coefficient of static friction of 1.2 and a coefficient of kinetic friction of 0.90.
5.116 For a person wearing these shoes, what’s the maximum
angle (with respect to the horizontal) of a smooth rock that can
be walked on without slipping? (a) 42°; (b) 50°; (c) 64°; (d) larger
than 90°.
5.117 If the person steps onto a smooth rock surface that’s inclined at an angle large enough that these shoes begin to slip, what
will happen? (a) She will slide a short distance and stop; (b) she
will accelerate down the surface; (c) she will slide down the surface at constant speed; (d) we can’t tell what will happen without
knowing her mass.
5.118 A person wearing these shoes stands on a smooth, horizontal rock. She pushes against the ground to begin running. What
is the maximum horizontal acceleration she can have without slipping? (a) 0.20g; (b) 0.75g; (c) 0.90g; (d) 1.2g.
answers
171
answers
chapter opening question
?
(iii) The upward force exerted by the air has the same magnitude as the force of gravity. Although the seed and pappus are
descending, their vertical velocity is constant, so their vertical
acceleration is zero. According to Newton’s first law, the net vertical force on the seed and pappus must also be zero. The individual vertical forces must balance.
test your Understanding questions
5.1 (ii) The two cables are arranged symmetrically, so the tension in either cable has the same magnitude T. The vertical component of the tension from each cable is T sin 45° (or, equivalently,
T cos 45°), so Newton’s first law applied to the vertical forces
tells us that 2T sin 45° - w = 0. Hence T = w>12 sin 45°2 =
w>12 = 0.71w. Each cable supports half of the weight of the
traffic light, but the tension is greater than w>2 because only the
vertical component of the tension counteracts the weight.
5.2 (ii) No matter what the instantaneous velocity of the glider,
its acceleration is constant and has the value found in Example 5.12.
In the same way, the acceleration of a body in free fall is the same
whether it is ascending, descending, or at the high point of its
motion (see Section 2.5).
5.3 (a): (i), (iii); (b): (ii), (iv); (c): (v) In situations (i) and (iii)
the box is not accelerating (so the net force on it must be zero)
and no other force is acting parallel to the horizontal surface;
hence no friction force is needed to prevent sliding. In situations
(ii) and (iv) the box would start to slide over the surface if no
friction were present, so a static friction force must act to prevent
this. In situation (v) the box is sliding over a rough surface, so a
kinetic friction force acts on it.
5.4 (iii) A satellite of mass m orbiting the earth at speed v in
an orbit of radius r has an acceleration of magnitude v 2>r, so
the net force acting on it from the earth’s gravity has magnitude
F = mv 2>r. The farther the satellite is from the earth, the greater
the value of r, the smaller the value of v, and hence the smaller
the values of v 2>r and of F. In other words, the earth’s gravitational force decreases with increasing distance.
bridging Problem
(a) Tmax = 2p
h1cos b + ms sin b2
B g tan b1sin b - ms cos b2
(b) Tmin = 2p
h1cos b - ms sin b2
B g tan b1sin b + ms cos b2
?
A baseball pitcher does
work with his throwing
arm to give the ball a property
called kinetic energy, which
depends on the ball’s mass and
speed. Which has the greatest
kinetic energy? (i) A ball of mass
0.145 kg moving at 20.0 m>s;
(ii) a smaller ball of mass
0.0145 kg moving at 200 m>s;
(iii) a larger ball of mass 1.45 kg
moving at 2.00 m>s; (iv) all
three balls have the same
kinetic energy; (v) it depends
on the directions in which the
balls move.
6
Work and
kinetic energy
Learning goaLs
Looking forward at …
6.1 What it means for a force to do work on a
body, and how to calculate the amount of
work done.
6.2 The definition of the kinetic energy (energy
of motion) of a body, and how the total work
done on a body changes the body’s kinetic
energy.
6.3 How to use the relationship between total
work and change in kinetic energy when
the forces are not constant, the body follows
a curved path, or both.
6.4 How to solve problems involving power (the
rate of doing work).
Looking back at …
1.10 The scalar product (or dot product) of two
vectors.
2.4 Straight-line motion with constant
acceleration.
4.3 Newton’s second law.
4.5 Newton’s third law.
5.1, 5.2 Using components to find the net force.
172
S
uppose you try to find the speed of an arrow that has been shot from a
bow. You apply Newton’s laws and all the problem-solving techniques
that we’ve learned, but you run across a major stumbling block: After the
archer releases the arrow, the bow string exerts a varying force that depends
on the arrow’s position. As a result, the simple methods that we’ve learned
aren’t enough to calculate the speed. Never fear; we aren’t by any means
finished with mechanics, and there are other methods for dealing with such
problems.
The new method that we’re about to introduce uses the ideas of work and energy.
The importance of the energy idea stems from the principle of conservation of
energy: Energy is a quantity that can be converted from one form to another but
cannot be created or destroyed. In an automobile engine, chemical energy stored
in the fuel is converted partially to the energy of the automobile’s motion and
partially to thermal energy. In a microwave oven, electromagnetic energy obtained from your power company is converted to thermal energy of the food being cooked. In these and all other processes, the total energy—the sum of all
energy present in all different forms—remains the same. No exception has ever
been found.
We’ll use the energy idea throughout the rest of this book to study a tremendous range of physical phenomena. This idea will help you understand how
automotive engines work, how a camera’s flash unit can produce a short burst
of light, and the meaning of Einstein’s famous equation E = mc2.
In this chapter, though, our concentration will be on mechanics. We’ll
learn about one important form of energy called kinetic energy, or energy of
motion, and how it relates to the concept of work. We’ll also consider power,
which is the time rate of doing work. In Chapter 7 we’ll expand these ideas
into a deeper understanding of the concepts of energy and the conservation
of energy.
173
6.1 Work
6.1 Work
6.1 These people are doing work as they
You’d probably agree that it’s hard work to pull a heavy sofa across the room, to
lift a stack of encyclopedias from the floor to a high shelf, or to push a stalled car
off the road. Indeed, all of these examples agree with the everyday meaning of
work—any activity that requires muscular or mental effort.
In physics, work has a much more precise definition. By making use of this
definition we’ll find that in any motion, no matter how complicated, the total
work done on a particle by all forces that act on it equals the change in its kinetic
energy—a quantity that’s related to the particle’s mass and speed. This relationship holds even when the forces acting on the particle aren’t constant, a situation
that can be difficult or impossible to handle with the techniques you learned in
Chapters 4 and 5. The ideas of work and kinetic energy enable us to solve problems in mechanics that we could not have attempted before.
In this section we’ll see how work is defined and how to calculate work in a
variety of situations involving constant forces. Later in this chapter we’ll relate
work and kinetic energy, and then apply these ideas to problems in which the
forces are not constant.
The three examples of work described above—pulling a sofa, lifting encyclopedias, and pushing a car—have something in common. In each case you do
work by exerting a force on a body while that body moves from one place to
another—that is, undergoes a displacement (Fig. 6.1). You do more work if the
force is greater (you push harder on the car) or if the displacement is greater (you
push the car farther down the road).
The physicist’s definition of work is based on these observations. Consider
a body that undergoes a displacement of magnitude s along a straight line. (For
now, we’ll assume that any body we discuss can be treated as a particle so that
we can ignore any rotation
or changes in shape of the body.) While the body
S
S
moves, a constant force F acts on it in the same direction as the displacement s
(Fig. 6.2). We define the work W done by this constant force under these
circumstances as the product of the force magnitude F and the displacement
magnitude s:
W = Fs
(constant force in direction of straight-line displacement)
(6.1)
The work done on the body is greater if either the force F or the displacement s is
greater, in agreement with our observations above.
Caution Work = W, weight = w Don’t confuse uppercase W (work) with lowercase
w (weight). Though the symbols are similar, work and weight are different quantities. ❙
The SI unit of work is the joule (abbreviated J, pronounced “jool,” and
named in honor of the 19th-century English physicist James Prescott Joule).
From Eq. (6.1) we see that in any system of units, the unit of work is the unit
of force multiplied by the unit of distance. In SI units the unit of force is
the newton and the unit of distance is the meter, so 1 joule is equivalent to
1 newton-meter 1N # m2:
1 joule = 11 newton211 meter2 or 1 J = 1 N # m
If you lift an object with a weight of 1 N (about the weight of a medium-sized
apple) a distance of 1 m at a constant speed, you exert a 1-N force on the object in
the same direction as its 1-m displacement and so do 1 J of work on it.
As an illustration of Eq. (6.1), think of a person pushing a Sstalled car. If he
S
pushes the car through a displacement s with a constant force F in the direction
of motion, the amount of work he does on the car is given by Eq. (6.1): W = Fs.
push on the car because they exert a force
on the car as it moves.
6.2 The work done by a constant force
acting in the same direction as the
displacement.
If a body moves through a
S
displacement s Swhile a
constant force F acts on it
in the same direction ...
S
F
x
S
s
... the work done by
the force on the
body is W = Fs.
BIO application Work and Muscle
Fibers Our ability to do work with our
bodies comes from our skeletal muscles. The
fiberlike cells of skeletal muscle, shown in this
micrograph, can shorten, causing the muscle
as a whole to contract and to exert force on
the tendons to which it attaches. Muscle
can exert a force of about 0.3 N per square
millimeter of cross-sectional area: The greater
the cross-sectional area, the more fibers the
muscle has and the more force it can exert
when it contracts.
174
Chapter 6 Work and Kinetic energy
6.3 The work done by a constant force acting at an angle to the displacement.
S
F
S
F
F# does no work on the car.
F# = F sin f
The car moves through
S
displacement s Swhile a
constant force F acts
on it at an angle f to
the displacement.
f
Only FΠdoes work on the car:
W = FŒs = (F cos f)s
= Fs cos f
FΠ= F cos f
S
s
But what
if the person pushes at an angle f to the car’s displacement (Fig. 6.3)?
S
S
Then F has a component FΠ= F cos f in the direction of the displacement s and
S
a component F# = F sin f that acts perpendicular to s . (Other
forces must act
S
S
on the car so that it moves along s , not in the direction of F. We’re interested
in only the work that the person does, however, so we’ll consider only the force
he exerts.) Only the parallel component FΠis effective in moving the car, so we
define the work as the product of this force component and the magnitude of the
displacement. Hence W = FŒs = 1F cos f2s, or
S
Work done on a particle
S
by constant force F during
straight-line displacement Ss
S
Magnitude of F
W = Fs cos f
S
S
Angle between F and s
(6.2)
S
Magnitude of s
S
If f = 0, so that F and s are in the same direction, then cos f = 1 and we are
back to Eq. (6.1).
Equation (6.2) has the form
of the scalar product of two vectors, which we
S
S
introduced in Section 1.10: A ~ B = AB cos f. You may want to review that definition. Hence we can write Eq. (6.2) more compactly as
Work done on a particle
S
by constant force F during
S
straight-line displacement s
S
S
W = F~s
(6.3)
S
S
Scalar product (dot product) of vectors F and s
Caution Work is a scalar An essential point: Work is a scalar quantity, even though
it’s calculated from two vector quantities (force and displacement). A 5-N force toward the
east acting on a body that moves 6 m to the east does the same amount of work as a 5-N
force toward the north acting on a body that moves 6 m to the north. ❙
Solution
ExamplE 6.1 Work donE by a constant forcE
(a) Steve exerts a steady force of magnitude 210 N (about 47 lb)
on the stalled car in Fig. 6.3 as he pushes it a distance of 18 m.
The car also has a flat tire, so to make the car track straight Steve
must push at an angle of 30° to the direction of motion. How much
work does Steve do? (b) In a helpful
mood, Steve pushes a second
S
stalled car with a steady force F = 1160 N2dn − 140 N2en . The disS
placement of the car is s = 114 m2dn + 111 m2en . How much work
does Steve do in this case?
soLution
identiFy and set up: In both parts (a) and (b), the target variable
is the work W done by Steve. In each case the force is constant and
the displacement is along a Sstraight line, so we can use Eq. (6.2)
S
or (6.3). The angle between F and s is given
in part (a), so we can
S
S
apply Eq. (6.2) directly. In part (b) both F and s are given in terms
of components,
so it’s best to calculate the scalar product by using
S
S
Eq. (1.19): A ~ B = AxBx + AyBy + AzBz .
exeCute: (a) From Eq. (6.2),
W = Fs cos f = 1210 N2118 m2cos 30° = 3.3 * 103 J
S
(b) The components of F are Fx = 160 N and Fy = - 40 N,
S
and the components of s are x = 14 m and y = 11 m. (There are
no z-components for either vector.) Hence, using Eqs. (1.19) and
(6.3), we have
S
S
W = F ~ s = Fx x + Fy y
= 1160 N2114 m2 + 1-40 N2111 m2
= 1.8 * 103 J
evaLuate: In each case the work that Steve does is more than
1000 J. This shows that 1 joule is a rather small amount of work.
175
6.1 Work
S
S
6.4 A constant force F can do positive, negative, or zero work depending on the angle between F and
S
the displacement s .
Direction of Force (or Force Component)
(a)
Situation
Force Diagram
S
S
F
S
Force F has a component in direction of displacement:
W = FŒs = (F cos f) s
Work is positive.
S
F
f
f
FΠ= F cosf
S
s
(b)
S
Force F has a component opposite to direction of displacement:
W = FŒs = (F cos f) s
Work is negative (because F cos f is negative for 90° 6 f 6 180°).
S
(c)
Force F (or force component F#) is perpendicular to direction
of displacement: The force (or force component) does no work
on the object.
S
S
F
F
F
f
F#
f
FΠ= F cos f
S
s
S
S
F
F#
S
F
F
S
F
f = 90°
S
s
Work: positive, negative, or Zero
In Example 6.1 the work done in pushing the cars was positive. But it’s important
to understand that work can also be negative or zero. This is the essential way in
which work as defined in physics differs from the “everyday” definition of work.
When the force has a component in the same direction as the displacement
(f between 0° and 90°), cos f in Eq. (6.2) is positive and the work W is positive
(Fig. 6.4a). When the force has a component opposite to the displacement
1f between 90° and 180°2, cos f is negative and the work is negative (Fig. 6.4b).
When the force is perpendicular to the displacement, f = 90° and the work
done by the force is zero (Fig. 6.4c). The cases of zero work and negative work
bear closer examination, so let’s look at some examples.
There are many situations in which forces act but do zero work. You might
think it’s “hard work” to hold a barbell motionless in the air for 5 minutes
(Fig. 6.5). But in fact, you aren’t doing any work on the barbell because there
is no displacement. (Holding the barbell requires you to keep the muscles of
your arms contracted, and this consumes energy stored in carbohydrates and
fat within your body. As these energy stores are used up, your muscles feel
fatigued even though you do no work on the barbell.) Even when you carry a
book while you walk with constant velocity on a level floor, you do no work
on the book. It has a displacement, but the (vertical) supporting force that you
exert on the book has no component in the direction of the (horizontal) motion.
Then f = 90° in Eq. (6.2), and cos f = 0. When a body slides along a surface,
the work done on the body by the normal force is zero; and when a ball on a
string moves in uniform circular motion, the work done on the ball by the tension
in the string is also zero. In both cases the work is zero because the force has
no component in the direction of motion.
What does it mean to do negative work? The answer comes from Newton’s
third law of motion. When a weightlifter lowers a barbell as in Fig. 6.6a (next
S
page), his hands and the barbell
move together with the same displacement s .
S
The barbell exerts a force Fbarbell on hands on his hands in the same direction as the
hands’ displacement, so the work done by the barbell on his hands is positive
(Fig. 6.6b). But Sby Newton’s third law
the weightlifter’s hands exert an equal and
S
opposite force Fhands on barbell = −Fbarbell on hands on the barbell (Fig. 6.6c). This
force, which keeps the barbell from crashing to the floor, acts opposite to the barbell’s displacement. Thus the work done by his hands on the barbell is negative.
Because the weightlifter’s hands and the barbell have the same displacement, the
6.5 A weightlifter does no work on a
barbell as long as he holds it stationary.
S
F
The weightlifter exerts
an upward force on the
barbell ...
... but because the
barbell is stationary (its
displacement is zero),
he does no work on it.
176
Chapter 6 Work and Kinetic energy
6.6 This weightlifter’s hands do negative work on a barbell as the barbell does positive work on his hands.
(a) A weightlifter lowers a barbell to the floor.
(b) The barbell does positive work on the
weightlifter’s hands.
(c) The weightlifter’s hands do negative work
on the barbell.
S
Fhands on barbell
S
s
S
Fbarbell on hands
The force of the barbell on the
weightlifter’s hands is in the same
direction as the hands’ displacement.
S
s
S
s
The force of the weightlifter’s hands
on the barbell is opposite to the
barbell’s displacement.
work that his hands do on the barbell is just the negative of the work that the barbell does on his hands. In general, when one body does negative work on a second
body, the second body does an equal amount of positive work on the first body.
Data SpeakS
Positive, negative, and Zero Work
When students were given a problem that
required them to find the work done by
a constant force during a straight-line
displacement, more than 59% gave an
incorrect answer. Common errors:
● Forgetting that a force does negative
work if it acts opposite to the direction
of the object’s displacement.
●
Forgetting that, even if a force is present,
it does zero work if it acts perpendicular
to the direction of the displacement.
total Work
How do we calculate work when several forces act on a body? One way is to use
Eq. (6.2) or (6.3) to compute the work done by each separate force. Then, because
work is a scalar quantity, the total work Wtot done on the body by all the forces
is the algebraic sum of the quantities of work done by the individual forces. An
alternative way to find the total work Wtot is to compute the
vector sum of the
S
forces (that is, the net force) and then use this vector sum as F in Eq. (6.2) or (6.3).
The following example illustrates both of these techniques.
Solution
ExamplE 6.2
Caution Keep track of who’s doing the work We always speak of work done on a particular body by a specific force. Always specify exactly what force is doing the work.
When you lift a book, you exert an upward force on it and the book’s displacement is
upward, so the work done by the lifting force on the book is positive. But the work done by
the gravitational force (weight) on a book being lifted is negative because the downward
gravitational force is opposite to the upward displacement. ❙
WorK donE by sEvEral forcEs
A farmer hitches her tractor to a sled loaded with firewood and
pulls it a distance of 20 m along level ground (Fig. 6.7a). The total
weight of sled and load is 14,700 N. The tractor exerts a constant
5000-N force at an angle of 36.9° above the horizontal. A 3500-N
friction force opposes the sled’s motion. Find the work done by each
force acting on the sled and the total work done by all the forces.
6.7 Calculating the work done on a sled of firewood being pulled
by a tractor.
(a)
(b) Free-body diagram for sled
f
Solution
iDentify and Set uP: Each force is constant and the sled’s dis-
placement is along a straight line, so we can use the ideas of this
section to calculate the work. We’ll find the total work in two
ways: (1) by adding the work done on the sled by each force and
(2) by finding the work done by the net force on the sled. We
first draw a free-body diagram showing all of the forces acting
on the sled, and we choose a coordinate system (Fig. 6.7b). For
each force—weight, normal force, force of the tractor, and friction force—we know the angle between the displacement (in the
positive x-direction) and the force. Hence we can use Eq. (6.2) to
calculate the work each force does.
As in Chapter 5, we’ll find the net force by adding the components of the four forces. Newton’s second law tells us that because
the sled’s motion is purely horizontal, the net force can have only
a horizontal component.
exeCute: (1) The work Ww done by the weight is zero because its
direction is perpendicular to the displacement (compare Fig. 6.4c).
For the same reason, the work Wn done by the normal force is also
zero. (Note that we don’t need to calculate the magnitude n to conclude this.) So Ww = Wn = 0.
6.2 Kinetic energy and the Work–energy theorem
177
gFx = FT cos f + 1- f2 = 15000 N2 cos 36.9° - 3500 N
That leaves the work WT done by the force FT exerted by the
tractor and the work Wf done by the friction force f. From Eq. (6.2),
= 500 N
gFy = FT sin f + n + 1- w2
WT = FT s cos 36.9° = 15000 N2120 m210.8002 = 80,000 N # m
= 80 kJ
= 15000 N2 sin 36.9° + n - 14,700 N
S
The friction force f is opposite to the displacement, so for this
force f = 180° and cos f = - 1. Again from Eq. (6.2),
We don’t need the second equation; we know that the y-component
of force is perpendicular to the displacement, so it does no work.
Besides, there is no y-component of acceleration, so gFy must
be zero anyway. The total work is therefore the work done by the
total x-component:
Wf = fs cos 180° = 13500 N2120 m21-12 = - 70,000 N # m
= - 70 kJ
The total work Wtot done on the sled by all forces is the algebraic
sum of the work done by the individual forces:
S
S
Wtot = 1gF2 ~ s = 1gFx2s = 1500 N2120 m2 = 10,000 J
= 10 kJ
Wtot = Ww + Wn + WT + Wf = 0 + 0 + 80 kJ + 1- 70 kJ2
= 10 kJ
evaLuate: We get the same result for Wtot with either method, as
(2) In the second approach, we first find the vector sum of
all the forces (the net force) and then use it to compute the total
work. It’s easiest to find the net force by using components.
From Fig. 6.7b,
we should. Note that the net force in the x-direction is not zero,
and so the sled must accelerate as it moves. In Section 6.2 we’ll
return to this example and see how to use the concept of work to
explore the sled’s changes of speed.
test your understanding oF seCtion 6.1 An electron moves in a straight
line toward the east with a constant speed of 8 * 107 m>s. It has electric, magnetic, and
gravitational forces acting on it. During a 1-m displacement, the total work done on the
electron is (i) positive; (ii) negative; (iii) zero; (iv) not enough information is given. ❙
6.2 kinetiC energy and the
Work–energy theoreM
The total work done on a body by external forces is related to the body’s
displacement—that is, to changes in its position. But the total work is also
related to changes in the speed of the body. To see this, consider Fig. 6.8, which
shows a block sliding on a frictionless table. The
forces acting on the block are
S
S
S
its weight w, the normal force n, and the force F exerted on it by the hand.
In Fig. 6.8a the net force on the block is in the direction of its motion. From
Newton’s second law, this means that the block speeds up; from Eq. (6.1), this
also means that the total work Wtot done on the block is positive. The total work
PhET: The Ramp
6.8 The relationship between the total work done on a body and how the body’s speed changes.
(a)
A block slides to the right on a frictionless surface.
v
(b)
(c)
v
If you push to the right
on the block as it moves,
the net force on the
block is to the right.
v
If you push to the left
on the block as it moves,
the net force on the
block is to the left.
If you push straight
down on the block as
it moves, the net force
on the block is zero.
n
n
n
S
s
F
w
• The total work done on the block during
S
a displacement s is positive: Wtot 7 0.
• The block speeds up.
S
S
s
s
F
w
• The total work done on the block during
S
a displacement s is negative: Wtot 6 0.
• The block slows down.
w
F
• The total work done on the block during
S
a displacement s is zero: Wtot = 0.
• The block’s speed stays the same.
178
Chapter 6 Work and Kinetic energy
S
6.9 A constant net force F does work on a
moving body.
Speed v1
m
Speed v2
S
Net force F
m
x
x1
S
s
x2
is negative in Fig. 6.8b because the net force opposes the displacement; in this
case the block slows down. The net force is zero in Fig. 6.8c, so the speed of the
block stays the same and the total work done on the block is zero. We can conclude that when a particle undergoes a displacement, it speeds up if Wtot 7 0,
slows down if Wtot 6 0, and maintains the same speed if Wtot = 0.
Let’s make this more quantitative. In Fig. 6.9 a particle with mass m moves
along the x-axis under the action of a constant net force with magnitude F that
points in the positive x-direction. The particle’s acceleration is constant and
given by Newton’s second law (Section 4.3): F = max. As the particle moves
from point x1 to x2, it undergoes a displacement s = x2 - x1 and its speed
changes from v1 to v2. Using a constant-acceleration equation from Section 2.4,
Eq. (2.13), and replacing v0x by v1, vx by v2, and 1x - x02 by s, we have
v 22 = v 12 + 2axs
v 22 - v 12
2s
When we multiply this equation by m and equate max to the net force F, we find
ax =
F = max = m
v 22 - v 12
2s
and
Fs = 12 mv 22 - 12 mv 12
(6.4)
In Eq. (6.4) the product Fs is the work done by the net force F and thus is equal
to the total work Wtot done by all the forces acting on the particle. The quantity
1
2
2 mv is called the kinetic energy K of the particle:
Kinetic energy
of a particle
K = 12 mv 2 of different bodies.
S
v
S
v
m
Same mass, same speed, different directions
of motion: same kinetic energy
m
2m
S
v
S
v
Twice the mass, same speed:
twice the kinetic energy
m
S
v
m
Mass of particle
(6.5)
Speed of particle
?
6.10 Comparing the kinetic energy
m
K = 12 mv2
S
2v
Same mass, twice the speed:
four times the kinetic energy
Like work, the kinetic energy of a particle is a scalar quantity; it depends on
only the particle’s mass and speed, not its direction of motion (Fig. 6.10).
Kinetic energy can never be negative, and it is zero only when the particle is
at rest.
We can now interpret Eq. (6.4) in terms of work and kinetic energy. The first
term on the right side of Eq. (6.4) is K2 = 12 mv 22, the final kinetic energy of the
particle (that is, after the displacement). The second term is the initial kinetic
energy, K1 = 12 mv 12, and the difference between these terms is the change in
kinetic energy. So Eq. (6.4) says:
Work–energy theorem: Work done by the net force on a particle
equals the change in the particle’s kinetic energy.
Total work done
on particle =
Wtot = K2
work done by
net force
Final kinetic energy
- K1 = ∆K
Change in
kinetic energy
(6.6)
Initial kinetic energy
This work–energy theorem agrees with our observations about the block in
Fig. 6.8. When Wtot is positive, the kinetic energy increases (the final kinetic
energy K2 is greater than the initial kinetic energy K1) and the particle is going
faster at the end of the displacement than at the beginning. When Wtot is negative, the kinetic energy decreases (K2 is less than K1) and the speed is less after
the displacement. When Wtot = 0, the kinetic energy stays the same 1K1 = K22
and the speed is unchanged. Note that the work–energy theorem by itself tells
us only about changes in speed, not velocity, since the kinetic energy doesn’t
depend on the direction of motion.
From Eq. (6.4) or Eq. (6.6), kinetic energy and work must have the same units.
Hence the joule is the SI unit of both work and kinetic energy (and, as we will see
6.2 Kinetic energy and the Work–energy theorem
179
later, of all kinds of energy). To verify this, note that in SI the quantity K = 12 mv 2
has units kg # 1m>s22 or kg # m2>s2; we recall that 1 N = 1 kg # m>s2, so
1 J = 1 N # m = 1 1kg # m>s22 # m = 1 kg # m2>s2
Because we used Newton’s laws in deriving the work–energy theorem, we can
use this theorem only in an inertial frame of reference. Note that the work–energy
theorem is valid in any inertial frame, but the values of Wtot and K2 - K1 may
differ from one inertial frame to another (because the displacement and speed of a
body may be different in different frames).
We’ve derived the work–energy theorem for the special case of straight-line
motion with constant forces, and in the following examples we’ll apply it to this
special case only. We’ll find in the next section that the theorem is valid even
when the forces are not constant and the particle’s trajectory is curved.
problEm-solving stratEgy 6.1
Work and kinEtic EnErgy
identiFy the relevant concepts: The work–energy theorem,
Wtot = K2 - K1, is extremely useful when you want to relate a
body’s speed v1 at one point in its motion to its speed v2 at a different point. (It’s less useful for problems that involve the time it
takes a body to go from point 1 to point 2 because the work–energy
theorem doesn’t involve time at all. For such problems it’s usually
best to use the relationships among time, position, velocity, and
acceleration described in Chapters 2 and 3.)
set up the problem using the following steps:
1. Identify the initial and final positions of the body, and draw a
free-body diagram showing all the forces that act on the body.
2. Choose a coordinate system. (If the motion is along a straight
line, it’s usually easiest to have both the initial and final positions lie along one of the axes.)
3. List the unknown and known quantities, and decide which
unknowns are your target variables. The target variable may
be the body’s initial or final speed, the magnitude of one of the
forces acting on the body, or the body’s displacement.
exeCute the solution: Calculate the work W done by each force.
If the force is constant and the displacement is a straight line, you
can use Eq. (6.2) or Eq. (6.3). (Later in this chapter we’ll see
how to handle varying forces and curved trajectories.) Be sure
to check signs; W must be positive if the force has a component
in the direction of the displacement, negative if the force has a
component opposite to the displacement, and zero if the force and
displacement are perpendicular.
Add the amounts of work done by each force to find the total
work Wtot . Sometimes it’s easier to calculate the vector sum of the
forces (the net force) and then find the work done by the net force;
this value is also equal to Wtot .
Write expressions for the initial and final kinetic energies, K1
and K2. Note that kinetic energy involves mass, not weight; if you
are given the body’s weight, use w = mg to find the mass.
Finally, use Eq. (6.6), Wtot = K2 - K1, and Eq. (6.5),
K = 12 mv 2, to solve for the target variable. Remember that the
right-hand side of Eq. (6.6) represents the change of the body’s
kinetic energy between points 1 and 2; that is, it is the final kinetic
energy minus the initial kinetic energy, never the other way
around. (If you can predict the sign of Wtot, you can predict
whether the body speeds up or slows down.)
evaLuate your answer: Check whether your answer makes sense.
Remember that kinetic energy K = 12 mv 2 can never be negative. If
you come up with a negative value of K, perhaps you interchanged
the initial and final kinetic energies in Wtot = K2 - K1 or made a
sign error in one of the work calculations.
Solution
ExamplE 6.3 Using Work and EnErgy to calcUlatE spEEd
Let’s look again at the sled in Fig. 6.7 and our results from
Example 6.2. Suppose the sled’s initial speed v1 is 2.0 m>s. What
is the speed of the sled after it moves 20 m?
6.11 Our sketch for this problem.
soLution
identiFy and set up: We’ll use the work–energy theorem,
Eq. (6.6), Wtot = K2 - K1, since we are given the initial speed
v1 = 2.0 m > s and want to find the final speed v2. Figure 6.11
shows our sketch of the situation. The motion is in the positive
x-direction. In Example 6.2 we calculated the total work done by
all the forces: Wtot = 10 kJ . Hence the kinetic energy of the sled
and its load must increase by 10 kJ, and the speed of the sled must
also increase.
exeCute: To write expressions for the initial and final kinetic
energies, we need the mass of the sled and load. The combined
weight is 14,700 N, so the mass is
m =
14,700 N
w
= 1500 kg
=
g
9.8 m>s2
Continued
180
Chapter 6 Work and Kinetic energy
use the equations of motion for constant acceleration to find v2.
Since the acceleration is along the x-axis,
Then the initial kinetic energy K1 is
K1 = 12 mv 12 = 1211500 kg212.0 m>s22 = 3000 kg # m2>s2
= 3000 J
gFx
500 N
= 0.333 m>s2
=
m
1500 kg
a = ax =
The final kinetic energy K2 is
Then, using Eq. (2.13),
K2 = 12 mv 22 = 1211500 kg2v 22
v 22 = v 12 + 2as = 12.0 m>s22 + 210.333 m>s22120 m2
The work–energy theorem, Eq. (6.6), gives
= 17.3 m2>s2
K2 = K1 + Wtot = 3000 J + 10,000 J = 13,000 J
Setting these two expressions for K2 equal, substituting 1 J =
1 kg # m2>s2, and solving for the final speed v2, we find
v2 = 4.2 m>s
evaLuate: The total work is positive, so the kinetic energy increases 1K2 7 K12 and the speed increases 1v2 7 v12.
This problem can also be solved without Sthe work–energy
S
theorem. We can find the acceleration from gF = ma and then
This is the same result we obtained with the work–energy approach, but there we avoided the intermediate step of finding the
acceleration. You will find several other examples in this chapter
and the next that can be done without using energy considerations
but that are easier when energy methods are used. When a problem can be done by two methods, doing it by both methods (as we
did here) is a good way to check your work.
Solution
ExamplE 6.4
v2 = 4.2 m>s
forcEs on a hammErhEad
The 200-kg steel hammerhead of a pile driver is lifted 3.00 m
above the top of a vertical I-beam being driven into the ground
(Fig. 6.12a). The hammerhead is then dropped, driving the
I-beam 7.4 cm deeper into the ground. The vertical guide rails
exert a constant 60-N friction force on the hammerhead. Use the
work–energy theorem to find (a) the speed of the hammerhead
just as it hits the I-beam and (b) the average force the hammerhead
exerts on the I-beam. Ignore the effects of the air.
soLution
identiFy: We’ll use the work–energy theorem to relate the ham-
merhead’s speed at different locations and the forces acting on it.
There are three locations of interest: point 1, where the hammerhead starts from rest; point 2, where it first contacts the I-beam;
and point 3, where the hammerhead and I-beam come to a halt
(Fig. 6.12a). The two target variables are the hammerhead’s speed
at point 2 and the average force the hammerhead exerts between
points 2 and 3. Hence we’ll apply the work–energy theorem twice:
once for the motion from 1 to 2, and once for the motion from 2 to 3.
set up: Figure 6.12b shows the vertical forces on the hammerhead as it falls from point 1 to point 2. (We can ignore any horizontal forces that may be present because they do no work as the
hammerhead moves vertically.) For this part of the motion, our
target variable is the hammerhead’s final speed v2.
6.12 (a) A pile driver pounds an I-beam into the ground. (b), (c) Free-body diagrams. Vector lengths
are not to scale.
(a)
(b) Free-body diagram
for falling hammerhead
(c) Free-body diagram for hammerhead
when pushing I-beam
y
y
Point 1
n
f = 60 N
x
3.00 m
v
w = mg
Point 2
7.4 cm
Point 3
f = 60 N
x
w = mg
6.2 Kinetic energy and the Work–energy theorem
Figure 6.12c shows the vertical forces on the hammerhead
during the motion from point 2 to point 3. In addition to the forces
shown in Fig. 6.12b, the I-beam exerts an upward normal force
of magnitude n on the hammerhead. This force actually varies as
the hammerhead comes to a halt, but for simplicity we’ll treat n as
a constant. Hence n represents the average value of this upward
force during the motion. Our target variable for this part of the
motion is the force that the hammerhead exerts on the I-beam; it is
the reaction force to the normal force exerted by the I-beam, so by
Newton’s third law its magnitude is also n.
exeCute: (a) From point 1 to point 2, the vertical forces are the
downward weight w = mg = 1200 kg219.8 m>s22 = 1960 N and
the upward friction force f = 60 N. Thus the net downward force
is w - f = 1900 N. The displacement of the hammerhead from
point 1 to point 2 is downward and equal to s12 = 3.00 m. The
total work done on the hammerhead between point 1 and point 2
is then
Wtot = 1w - f 2s12 = 11900 N213.00 m2 = 5700 J
At point 1 the hammerhead is at rest, so its initial kinetic energy
K1 is zero. Hence the kinetic energy K2 at point 2 equals the total
work done on the hammerhead between points 1 and 2:
Wtot = K2 - K1 = K2 - 0 = 12 mv 22 - 0
v2 =
215700 J2
2Wtot
= 7.55 m>s
=
B m
B 200 kg
This is the hammerhead’s speed at point 2, just as it hits the I-beam.
181
(b) As the hammerhead moves downward from point 2 to
point 3, its displacement is s23 = 7.4 cm = 0.074 m and the net
downward force acting on it is w - f - n (Fig. 6.12c). The total
work done on the hammerhead during this displacement is
Wtot = 1w - f - n2s23
The initial kinetic energy for this part of the motion is K2, which
from part (a) equals 5700 J. The final kinetic energy is K3 = 0
(the hammerhead ends at rest). From the work–energy theorem,
Wtot = 1w - f - n2s23 = K3 - K2
n = w - f -
K3 - K2
s23
= 1960 N - 60 N -
0 J - 5700 J
= 79,000 N
0.074 m
The downward force that the hammerhead exerts on the I-beam
has this same magnitude, 79,000 N (about 9 tons)—more than
40 times the weight of the hammerhead.
evaLuate: The net change in the hammerhead’s kinetic energy
from point 1 to point 3 is zero; a relatively small net force does
positive work over a large distance, and then a much larger net
force does negative work over a much smaller distance. The same
thing happens if you speed up your car gradually and then drive it
into a brick wall. The very large force needed to reduce the kinetic
energy to zero over a short distance is what does the damage to
your car—and possibly to you.
the Meaning of kinetic energy
Example 6.4 gives insight into the physical meaning of kinetic energy. The hammerhead is dropped from rest, and its kinetic energy when it hits the I-beam
equals the total work done on it up to that point by the net force. This result is
true in general: To accelerate a particle of mass m from rest (zero kinetic energy)
up to a speed v, the total work done on it must equal the change in kinetic energy
from zero to K = 12 mv 2:
Wtot = K - 0 = K
So the kinetic energy of a particle is equal to the total work that was done to accelerate it from rest to its present speed (Fig. 6.13). The definition K = 12 mv 2,
Eq. (6.5), wasn’t chosen at random; it’s the only definition that agrees with this
interpretation of kinetic energy.
In the second part of Example 6.4 the kinetic energy of the hammerhead
did work on the I-beam and drove it into the ground. This gives us another
interpretation of kinetic energy: The kinetic energy of a particle is equal to
the total work that particle can do in the process of being brought to rest.
This is why you pull your hand and arm backward when you catch a ball. As
the ball comes to rest, it does an amount of work (force times distance) on
your hand equal to the ball’s initial kinetic energy. By pulling your hand back,
you maximize the distance over which the force acts and so minimize the
force on your hand.
6.13 Imparting kinetic energy to a cue ball.
When a billiards player hits a cue ball at rest,
the ball’s kinetic energy after being hit is equal
to the work that was done on it by the cue.
The greater the force exerted by the cue and
the greater the distance the ball moves while
in contact with it, the greater the ball’s
kinetic energy.
182
Chapter 6 Work and Kinetic energy
comparing kinEtic EnErgiEs
Two iceboats like the one in Example 5.6 (Section 5.2) hold a race
on a frictionless horizontal lake (Fig. 6.14). The two iceboats have
masses m and 2m. The iceboatsS have identical sails, so the wind
exerts the same constant force F on each iceboat. They start from
rest and cross the finish line a distance s away. Which iceboat
crosses the finish line with greater kinetic energy?
soLution
If you use the definition of kinetic energy, K = 12 mv 2, Eq. (6.5),
the answer to this problem isn’t obvious. The iceboat of mass 2m
6.14 A race between iceboats.
F
F
m
2m
s
Start
Finish
Solution
concEptUal ExamplE 6.5
has greater mass, so you might guess that it has greater kinetic
energy at the finish line. But the lighter iceboat, of mass m, has
greater acceleration and crosses the finish line with a greater
speed, so you might guess that this iceboat has the greater kinetic
energy. How can we decide?
The key is to remember that the kinetic energy of a particle is
equal to the total work done to accelerate it from rest. Both iceboats travel the same distance s from rest, and only the horizontal
force F in the direction of motion does work on either iceboat.
Hence the total work done between the starting line and the finish
line is the same for each iceboat, Wtot = Fs. At the finish line,
each iceboat has a kinetic energy equal to the work Wtot done on it,
because each iceboat started from rest. So both iceboats have the
same kinetic energy at the finish line!
You might think this is a “trick” question, but it isn’t. If you
really understand the meanings of quantities such as kinetic energy,
you can solve problems more easily and with better insight.
Notice that we didn’t need to know anything about how much
time each iceboat took to reach the finish line. This is because the
work–energy theorem makes no direct reference to time, only to
displacement. In fact the iceboat of mass m has greater acceleration and so takes less time to reach the finish line than does the
iceboat of mass 2m.
Work and kinetic energy in Composite systems
6.15 The external forces acting on a
skater pushing off a wall. The work done
by these forces is zero, but the skater’s
kinetic energy changes nonetheless.
S
F
S
w
S
n1
S
n2
In this section we’ve been careful to apply the work–energy theorem only to
bodies that we can represent as particles—that is, as moving point masses. New
subtleties appear for more complex systems that have to be represented as many
particles with different motions. We can’t go into these subtleties in detail in this
chapter, but here’s an example.
Suppose a boy stands on frictionless roller skates on a level surface, facing a
rigid wall (Fig. 6.15). He pushes against the wall, which makes him move to the
S
S
right. The forces acting on him are his weight w, the upward normal
forces n1
S
S
and n2 exerted by the ground on his skates, and the horizontal force F exerted on
S S
S
him bySthe wall. There is no vertical displacement, so w, n1, and n2 do no work.
Force F accelerates him to the right, but the parts of his body where that force
is
S
applied (the boy’s hands) do not move while the force acts. Thus the force F also
does no work. Where, then, does the boy’s kinetic energy come from?
The explanation is that it’s not adequate to represent the boy as a single point
mass. Different parts of the boy’s body have different motions; his hands remain
stationary against the wall while his torso is moving away from the wall. The various parts of his body interact with each other, and one part can exert forces and do
work on another part. Therefore the total kinetic energy of this composite system
of body parts can change, even though no work is done by forces applied by bodies
(such as the wall) that are outside the system. In Chapter 8 we’ll consider further
the motion of a collection of particles that interact with each other. We’ll discover
that just as for the boy in this example, the total kinetic energy of such a system can
change even when no work is done on any part of the system by anything outside it.
test your understanding oF seCtion 6.2 Rank the following bodies in
order of their kinetic energy, from least to greatest. (i) A 2.0-kg body moving at 5.0 m>s;
(ii) a 1.0-kg body that initially was at rest and then had 30 J of work done on it; (iii) a 1.0-kg
body that initially was moving at 4.0 m>s and then had 20 J of work done on it; (iv) a 2.0-kg
body that initially was moving at 10 m>s and then did 80 J of work on another body. ❙
183
6.3 Work and energy with Varying Forces
6.3 Work and energy With varying ForCes
So far we’ve considered work done by constant forces only. But what happens
when you stretch a spring? The more you stretch it, the harder you have to pull,
so the force you exert is not constant as the spring is stretched. We’ve also
restricted our discussion to straight-line motion. There are many situations in
which a body moves along a curved path and is acted on by a force that varies in magnitude, direction, or both. We need to be able to compute the work
done by the force in these more general cases. Fortunately, the work–energy
theorem holds true even when forces are varying and when the body’s path is
not straight.
6.16 Calculating the work done by a
varying force Fx in the x-direction as a
particle moves from x1 to x2.
(a) A particle moves from x1 to x2 in response to
a changing force in the x-direction.
F1x
F2x
x1
(b) The force Fx varies with position x ...
Fx
F2x
Graph of force
as a function
of position
Work done by a varying Force, straight-Line Motion
To add only one complication at a time, let’s consider straight-line motion along
the x-axis with a force whose x-component Fx may change as the body moves.
(A real-life example is driving a car along a straight road with stop signs, so the
driver has to alternately step on the gas and apply the brakes.) Suppose a particle
moves along the x-axis from point x1 to x2 (Fig. 6.16a). Figure 6.16b is a graph of
the x-component of force as a function of the particle’s coordinate x. To find the
work done by this force, we divide the total displacement into narrow segments
∆xa, ∆xb, and so on (Fig. 6.16c). We approximate the work done by the force during segment ∆xa as the average x-component of force Fax in that segment multiplied by the x-displacement ∆xa. We do this for each segment and then add the
results for all the segments. The work done by the force in the total displacement
from x1 to x2 is approximately
W = Fax ∆xa + Fbx ∆xb + g
F1x
O
x1
x
x2
x2 - x1
(c) ... but over a short displacement ∆x, the force
is essentially constant.
Fx The height of each strip
Ff x
represents the average Fex
force for that
Fdx
interval.
Fcx
Fbx
Fax
O x1 ∆xa
In the limit that the number of segments becomes very large and the width of
each becomes very small, this sum becomes the integral of Fx from x1 to x2 :
Work done on a particle by a
varying x-component of
force Fx during straight-line
displacement along x-axis
x
x2
∆xb
∆xc
∆xd
∆xe
∆xf
x2
x
Upper limit = final position
W =
x2
Lx1
Fx dx
Integral of x-component
of force
(6.7)
Lower limit = initial position
PhET: Molecular Motors
PhET: Stretching DNA
Note that Fax ∆xa represents the area of the first vertical strip in Fig. 6.16c and
that the integral in Eq. (6.7) represents the area under the curve of Fig. 6.16b
between x1 and x2. On such a graph of force as a function of position, the total
work done by the force is represented by the area under the curve between the
initial and final positions. Alternatively, the work W equals the average force
that acts over the entire displacement, multiplied by the displacement.
In the special case that Fx , the x-component of the force, is constant, we can
take it outside the integral in Eq. (6.7):
x2
W =
Lx1
6.17 The work done by a constant force F
in the x-direction as a particle moves from
x1 to x2.
The rectangular area under the
graph represents the work done by
the constant force of magnitude F
Fx during displacement s:
W = Fs
x2
Fx dx = Fx
Lx1
dx = Fx1x2 - x12
(constant force)
But x2 - x1 = s, the total displacement of the particle. So in the case of a
constant force F, Eq. (6.7) says that W = Fs, in agreement with Eq. (6.1). The
interpretation of work as the area under the curve of Fx as a function of x also
holds for a constant force: W = Fs is the area of a rectangle of height F and
width s (Fig. 6.17).
Now let’s apply these ideas to the stretched spring. To keep a spring stretched
beyond its unstretched length by an amount x, we have to apply a force of equal
F
O
x1
x2
s = x2 - x1
x
184
Chapter 6 Work and Kinetic energy
6.18 The force needed to stretch an ideal
spring is proportional to the spring’s elongation: Fx = kx.
x
- Fx
Fx = kx
6.19 Calculating the work done to stretch
a spring by a length X.
The area under the graph represents the work
done on the spring as the spring is stretched
from x = 0 to a maximum value X:
1
W = kX2
2
Fx
magnitude at each end (Fig. 6.18). If the elongation x is not too great, the force
we apply to the right-hand end has an x-component directly proportional to x:
Fx = kx
W =
L0
X
Fx dx =
L0
X
kx dx = 12 kX 2
(6.9)
We can also obtain this result graphically. The area of the shaded triangle in
Fig. 6.19, representing the total work done by the force, is equal to half the product
of the base and altitude, or
kX
X
(6.8)
where k is a constant called the force constant (or spring constant) of the spring.
The units of k are force divided by distance: N>m in SI units. A floppy toy spring
such as a Slinky™ has a force constant of about 1 N>m; for the much stiffer
springs in an automobile’s suspension, k is about 105 N>m. The observation
that force is directly proportional to elongation for elongations that are not too
great was made by Robert Hooke in 1678 and is known as Hooke’s law. It really
shouldn’t be called a “law,” since it’s a statement about a specific device and not
a fundamental law of nature. Real springs don’t always obey Eq. (6.8) precisely,
but it’s still a useful idealized model. We’ll discuss Hooke’s law more fully in
Chapter 11.
To stretch a spring, we must do work. We apply equal and opposite forces to
the ends of the spring and gradually increase the forces. We hold the left end stationary, so the force we apply at this end does no work. The force at the moving
end does do work. Figure 6.19 is a graph of Fx as a function of x, the elongation
of the spring. The work done by this force when the elongation goes from zero to
a maximum value X is
Fx = k x
O
(force required to stretch a spring)
x
W = 121X21kX2 = 12 kX 2
This equation also says that the work is the average force kX>2 multiplied by the
total displacement X. We see that the total work is proportional to the square of
the final elongation X. To stretch an ideal spring by 2 cm, you must do four times
as much work as is needed to stretch it by 1 cm.
Equation (6.9) assumes that the spring was originally unstretched. If initially
the spring is already stretched a distance x1, the work we must do to stretch it to a
greater elongation x2 (Fig. 6.20a) is
x2
W =
Lx1
x2
Fx dx =
Lx1
2 elongation x1
kx(a)
dxStretching
= 12 kx 2x2a spring
- 12 kxfrom
(6.10)
1
to elongation
2
Use your knowledge of geometry to convince yourself that the trapezoidal area
under the graph in Fig. 6.20b is given by the expression in Eq. (6.10).
x
x = 0
x = x1
x = x2
6.20 Calculating the work done to stretch a spring from one elongation to a greater one.
(b) Force-versus-distance graph
(a) Stretching a spring from elongation x1
to elongation x 2
The trapezoidal area under the graph represents
the work done on the spring to stretch it from
1
1
x = x1 to x = x 2: W = 2 kx 22 - 2 kx 12.
x
x = 0
x = x1
Fx
kx2
x = x2
k x1
(b) Force-versus-distance graph
The trapezoidal area under the graph represents
the work done on the spring to stretch it from
1
1
x = x1 to x = x 2: W = 2 k x 22 - 2 k x 12.
Fx
x = 0
x = x1
x = x2
x
6.3 Work and energy with Varying Forces
185
If the spring has spaces between the coils when it is unstretched, then it can
also be compressed, and Hooke’s law holds for compression as well as stretching. In this case the force and displacement are in the opposite directions from
those shown in Fig. 6.18, so both Fx and x in Eq. (6.8) are negative. Since both
Fx and x are reversed, the force again is in the same direction as the displacement, and the work done by Fx is again positive. So the total work is still given
by Eq. (6.9) or (6.10), even when X is negative or either or both of x1 and x2 are
negative.
Caution Work done on a spring vs. work done by a spring Equation (6.10) gives the
work that you must do on a spring to change its length. If you stretch a spring that’s
originally relaxed, then x1 = 0, x2 7 0, and W 7 0: The force you apply to one end of
the spring is in the same direction as the displacement, and the work you do is positive.
By contrast, the work that the spring does on whatever it’s attached to is given by the
negative of Eq. (6.10). Thus, as you pull on the spring, the spring does negative work
on you. ❙
Solution
ExamplE 6.6 Work donE on a spring scalE
A woman weighing 600 N steps on a bathroom scale that contains
a stiff spring (Fig. 6.21). In equilibrium, the spring is compressed
1.0 cm under her weight. Find the force constant of the spring and
the total work done on it during the compression.
soLution
identiFy and set up: In equilibrium the upward force exerted
by the spring balances the downward force of the woman’s weight.
We’ll use this principle and Eq. (6.8) to determine the force
constant k, and we’ll use Eq. (6.10) to calculate the work W that the
woman does on the spring to compress it. We take positive values
of x to correspond to elongation (upward in Fig. 6.21), so that both
the displacement of the end of the spring (x) and the x-component
of the force that the woman exerts on it 1Fx2 are negative. The applied force and the displacement are in the same direction, so the
work done on the spring will be positive.
exeCute: The top of the spring is displaced by x = - 1.0 cm =
-0.010 m, and the woman exerts a force Fx = - 600 N on the
spring. From Eq. (6.8) the force constant is then
k =
6.21 Compressing a spring in a bathroom scale.
Because of our choice of axis, both the
force component and displacement are
negative. The work on the spring is positive.
Then, using x1 = 0 and x2 = - 0.010 m in Eq. (6.10), we have
W = 12 kx 22 - 12 kx 12
+x
Fx 6 0
- 1.0 cm
Fx
-600 N
=
= 6.0 * 104 N>m
x
-0.010 m
= 12 16.0 * 104 N>m21- 0.010 m22 - 0 = 3.0 J
evaLuate: The work done is positive, as expected. Our arbitrary
choice of the positive direction has no effect on the answer for W.
You can test this by taking the positive x-direction to be downward, corresponding to compression. Do you get the same values
for k and W as we found here?
Work–energy theorem for straight-Line Motion,
varying Forces
In Section 6.2 we derived the work–energy theorem, Wtot = K2 - K1, for the special case of straight-line motion with a constant net force. We can now prove that
this theorem is true even when the force varies with position. As in Section 6.2,
let’s consider a particle that undergoes a displacement x while being acted on by
a net force with x-component Fx, which we now allow to vary. Just as in Fig. 6.16,
we divide the total displacement x into a large number of small segments ∆x. We
can apply the work–energy theorem, Eq. (6.6), to each segment because the value
186
Chapter 6 Work and Kinetic energy
BIO application tendons are
nonideal springs Muscles exert forces
via the tendons that attach them to bones. A
tendon consists of long, stiff, elastic collagen
fibers. The graph shows how the tendon from
the hind leg of a wallaby (a small kangaroolike marsupial) stretches in response to an
applied force. The tendon does not exhibit
the simple, straight-line behavior of an ideal
spring, so the work it does has to be found by
integration [Eq. (6.7)]. The tendon exerts less
force while relaxing than while stretching. As a
result, the relaxing tendon does only about
93% of the work that was done to stretch it.
of Fx in each small segment is approximately constant. The change in kinetic
energy in segment ∆xa is equal to the work Fax ∆xa, and so on. The total change
of kinetic energy is the sum of the changes in the individual segments, and thus
is equal to the total work done on the particle during the entire displacement. So
Wtot = ∆K holds for varying forces as well as for constant ones.
Here’s an alternative derivation of the work–energy theorem for a force that
may vary with position. It involves making a change of variable from x to vx
in the work integral. Note first that the acceleration a of the particle can be
expressed in various ways, using ax = dvx>dt, vx = dx>dt, and the chain rule for
derivatives:
ax =
dvx
dvx dx
dvx
=
= vx
dt
dx dt
dx
(6.11)
From this result, Eq. (6.7) tells us that the total work done by the net force Fx is
x2
Wtot =
Lx1
Maximum tendon extension
Lx1
Lv1
Tendon relaxing
1
2
Extension (mm)
3
Solution
An air-track glider of mass 0.100 kg is attached to the end of a
horizontal air track by a spring with force constant 20.0 N>m
(Fig. 6.22a). Initially the spring is unstretched and the glider is
6.22 (a) A glider attached to an air track by a spring.
(b), (c) Our free-body diagrams.
(b) No friction
(a)
v1
(6.13)
This is the same as Eq. (6.6), so the work–energy theorem is valid even without
the assumption that the net force is constant.
ExamplE 6.7 motion With a varying forcE
k
(6.12)
mvx dvx
Wtot = 12 mv 22 - 12 mv 12
m
dvx
dx
dx
The integral of vx dvx is just v x2 >2. Substituting the upper and lower limits, we
finally find
Tendon being stretched
0
mvx
v2
Wtot =
500
x2
max dx =
Now 1dvx>dx2dx is the change in velocity dvx during the displacement dx, so
we can make that substitution in Eq. (6.12). This changes the integration variable from x to vx, so we change the limits from x1 and x2 to the corresponding
x-velocities v1 and v2:
Force exerted
by tendon (N)
1000
Lx1
x2
Fx dx =
(c) With friction
moving at 1.50 m>s to the right. Find the maximum distance d
that the glider moves to the right (a) if the air track is turned on,
so that there is no friction, and (b) if the air is turned off, so that
there is kinetic friction with coefficient mk = 0.47.
soLution
identiFy and set up: The force exerted by the spring is not
constant, so we cannot use the constant-acceleration formulas
of Chapter 2 to solve this problem. Instead, we’ll use the work–
energy theorem, since the total work done involves the distance
moved (our target variable). In Figs. 6.22b and 6.22c we choose
the positive x-direction to be to the right (in the direction of the
glider’s motion). We take x = 0 at the glider’s initial position
(where the spring is unstretched) and x = d (the target variable)
at the position where the glider stops. The motion is purely horizontal, so only the horizontal forces do work. Note that Eq. (6.10)
gives the work done by the glider on the spring as it stretches;
to use the work–energy theorem we need the work done by the
187
6.3 Work and Energy with Varying Forces
spring on the glider, which is the negative of Eq. (6.10). We expect
the glider to move farther without friction than with friction.
This is a quadratic equation for d. The solutions are
d = -
ExEcutE: (a) Equation (6.10) says that as the glider moves from
x1 = 0 to x2 = d, it does an amount of work W = 12 kd 2 - 12 k 1022 =
on the spring. The amount of work that the spring does on the
glider is the negative of this, - 12 kd 2. The spring stretches until the
glider comes instantaneously to rest, so the final kinetic energy K2
is zero. The initial kinetic energy is 12 mv 12, where v1 = 1.50 m>s
is the glider’s initial speed. From the work–energy theorem,
1
2
2 kd
We have
10.47210.100 kg219.80 m>s22
mkmg
= 0.02303 m
=
k
20.0 N>m
10.100 kg211.50 m>s22
mv 12
=
= 0.01125 m2
k
20.0 N>m
- 12 kd 2 = 0 - 12 mv 12
We solve for the distance d the glider moves:
d = v1
so
0.100 kg
m
= 11.50 m>s2
Ak
B 20.0 N>m
d = -10.02303 m2 { 210.02303 m22 + 0.01125 m2
= 0.106 m = 10.6 cm
= 0.086 m or
The stretched spring subsequently pulls the glider back to the left,
so the glider is at rest only instantaneously.
(b) If the air is turned off, we must include the work done by
the kinetic friction force. The normal force n is equal in magnitude to the weight of the glider, since the track is horizontal and
there are no other vertical forces. Hence the kinetic friction force
has constant magnitude fk = mkn = mkmg. The friction force is
directed opposite to the displacement, so the work done by friction is
Wfric = fkd cos 180° = - fkd = -mkmgd
The total work is the sum of Wfric and the work done by the spring,
- 12 kd 2. The work–energy theorem then says that
-mkmgd - 12 kd 2 = 0 - 12 mv 12
1
2
2 kd
+ mkmgd -
1
2
2 mv 1
EvaluatE: If we set mk = 0, our algebraic solution for d in part (b)
reduces to d = v1 1m>k, the zero-friction result from part (a).
With friction, the glider goes a shorter distance. Again the glider
stops instantaneously, and again the spring force pulls it toward
the left; whether it moves or not depends on how great the static
friction force is. How large would the coefficient of static friction ms have to be to keep the glider from springing back to
the left?
or
= 0
We can generalize our definition of work further to include a force that varies in
direction as well as magnitude, and a displacement that lies along a curved path.
Figure 6.23a shows a particle moving from P1 to P2 along a curve. We divide the
curve between these points into
manySinfinitesimal vector displacements, and we
S
S
call a typical one of these dl . Each dl is tangent to the path at its position. Let F
S
be theS force at a typical point along the path, and let f be the angle between F
and dl at this point. Then
the small element of work dW done on the particle durS
ing the displacement dl may be written as
S
dW = F ~ dl = F cos f dl = FΠdl
S
S
where FΠ= F cos f is the component
of F in the direction parallel to dl
S
(Fig. 6.23b). The work done by F on the particle as it moves from P1 to P2 is
S
6.23 A particle moves along a curved path
S
from point P1 to P2, acted on by a force F
that varies in magnitude and direction.
(a)
P2
S
F
P1
f
S
dl
S
During anSinfinitesimal displacement dl,
the force F does work dW on the particle:
S
S
dW = F ~ dl = F cos f dl
(b)
S
Scalar product (dot product) of F and displacement dl
Upper limit =
final position
W
Work done on
a particle by a S
varying force F
along a curved path
-0.132 m
The quantity d is a positive displacement, so only the positive
value of d makes sense. Thus with friction the glider moves a distance d = 0.086 m = 8.6 cm.
Work–Energy theorem for Motion along a curve
S
mkmg
mkmg 2
mv 12
a
b +
{
k
k
k
B
=
P2
LP1
S
S
F ~ dl =
Lower limit =
initial position
P2
LP1
F cos f dl =
LP1
Angle
between
S
S
F and dl
P2
S
P2
FΠdl
(6.14)
P1
F
F#
f
S
ComponentSof F
parallel to dl
S
dl
F‘ = F cos f
S
The integral in Eq. (6.14) (shown in three versions) is called a line integral. We’ll
see shortly how to evaluate an integral of this kind.
Only the component of F parallel to the
displacement, F‘ = FS cos f, contributes
to the work done by F.
188
Chapter 6 Work and Kinetic energy
We can now show that the work–energy theorem, Eq. (6.6), holds true
even
S
with varying forces and a displacement along a curved path.
The
force
F
is
esS
sentially constant over any given infinitesimal segment dl of the path, so we can
apply the work–energy theorem for straight-line motion to that segment. Thus
the change in the
particle’s
kinetic energy K over that segment equals the work
S
S
dW = FΠdl = F ~ dl done on the particle. Adding up these infinitesimal quantities of work from all the segments along the whole path gives the total work
done, Eq. (6.14), which equals the total change in kinetic energy over the whole
path. So Wtot = ∆K = K2 - K1 is true in general, no matter what the path and
no matter what the character of the forces. This can be proved more rigorously by
using steps like those in Eqs. (6.11) through (6.13).
Note that only the component of the net force parallel to the path, FŒ, does
work on the particle, so only this component can change the speed and kinetic
energy of the particle. The component perpendicular to the path, F# = F sin f,
has no effect on the particle’s speed; it acts only to change the particle’s direction.
To evaluate the line integral in Eq. (6.14) in a specific problem,
we need some
S
sort of detailed description of the path and of the way in which F varies along the
path. We usually express the line integral in terms of some scalar variable, as in
the following example.
Solution
ExamplE 6.8 motion on a curvEd path
At a family picnic you are appointed to push your obnoxious
cousin Throckmorton in a swing (Fig. 6.24a). His weight is w, the
length of the chains is R, and you push Throcky until the chains
make an angle u0Swith the vertical. To do this, you exert a varying
horizontal force F that starts at zero and gradually increases just
enough that Throcky and the swing move very slowly and remain
very nearly in equilibrium throughout the process. (a) What is the
total work done on Throcky by all forces? (b) What is the work
done by the tension
T in the chains? (c) What is the work you do
S
by exerting force F ? (Ignore the weight of the chains and seat.)
Solution
iDentify and Set uP: The motion is along a curve, so we’ll
use Eq. (6.14) to calculate the work
done by the net force, by the
S
tension force, and by the force F. Figure 6.24b shows our freebody diagram and coordinate system for some arbitrary point in
Throcky’s motion. We have replaced the sum of the tensions in the
two chains with a single tension T.
6.24 (a) Pushing cousin Throckmorton in a swing. (b) Our
free-body diagram.
(a)
(b) Free-body diagram for
Throckmorton (neglecting the
weight of the chains and seat)
u
R
S
S
F
s
dl
exeCute: (a) There are two ways to find the total work done dur-
ing the motion: (1) by calculating the work done by each force and
then adding those quantities, and (2) by calculating the work done
by the net force. The second approach is far easier here because
Throcky is nearly in equilibrium at every point. Hence the net
force on him is zero, the integral of the net force in Eq. (6.14) is
zero, and the total work done on him is zero.
(b) It’s also easy to find the work done by the chain tension T
because this force is perpendicular to the direction of motion at
all points along the path. Hence at all pointsSthe angle between the
chain tension and the displacement vector dl is 90° and the scalar
product in Eq. (6.14) is zero. Thus the chain
tension does zero work.
S
(c) To compute the work done by F, we need to calculate
the line integral in Eq. (6.14). Inside the integral is the quantity
F cos f dl; let’s see how to express each term inSthis quantity.
S
Figure 6.24a shows that the angle between F and dl is u, so we
replace f in Eq. (6.14) with u. The value of u changes as Throcky
moves.
S
To find the magnitude F of force F, note that the net force
on Throcky is zero (he is nearly in equilibrium at all points), so
gFx = 0 and gFy = 0. From Fig. 6.24b,
gFx = F + 1- T sin u2 = 0
gFy = T cos u + 1-w2 = 0
If you eliminate T from these two equations, you can show that
F = w tan u. As the angle u increases, the tangent increases and F
increases (you have to push harder).
S
To find the magnitude dl of the infinitesimal displacement dl ,
note that Throcky moves through a circular arc of radius R
(Fig. 6.24a). The arc length s equals the radius R multiplied byS
the length u (in radians): s = Ru. Therefore the displacement dl
corresponding to a small change of angle du has a magnitude
dl = ds = R du.
When we put all the pieces together, the integral in Eq. (6.14)
becomes
u
P2
W =
LP1
F cos f dl =
L0
u0
1w tan u2cos u1R du2 =
L0
u0
wR sin u du
6.4 power
(Recall that tan u = sin u>cos u, so tan u cos u = sin u.) We’ve converted the line integral into an ordinary integral in terms of the
angle u. The limits of integration are from the starting position at
u = 0 to the final position at u = u0. The final result is
W = wR
L0
u0
sin u du = - wR cos u 0 00 = -wR1cos u0 - 12
u
= wR11 - cos u02
y-components. The force of gravity has zero x-component
and a
S
y-component of -w. Figure 6.24a shows that dl has a magnitude
of ds, an x-component of ds cos u, and a y-component of ds sin u .
So
S
w = ne 1- w2
S
dl = nd 1ds cos u2 + ne 1ds sin u2
S
and W = 0, as we should expect. As u0 increases, cos u0 decreases
and W = wR11 - cos u02 increases. So the farther along the arc
you push Throcky, the more work you do. You can confirm that
the quantity R11 - cos u02 is equal to h, the increase in Throcky’s
height during the displacement. So the work that you do to raise
Throcky is just equal to his weight multiplied by the height that
you raise him.
We can check our results by calculating the work done by the
S
force of gravity w. From part (a) the total work done on Throcky
is zero, and from part (b) the work done by tension is zero. So
gravity must do a negative amount
of work that just balances the
S
positive work done by the force F that we calculated in part (c).
For variety, let’s calculate the work done by gravity
by using
S
S
the form of Eq. (6.14) that involves the
quantity
F
~
dl
,
and exS
S
press the force w and displacement dl in terms of their x- and
S
Use Eq. (1.19) to calculate the scalar product w ~ dl :
S
S
evaluate: If u0 = 0, there is no displacement; then cos u0 = 1
189
w ~ dl = 1- w21ds sin u2 = - w sin u ds
Using ds = R du, we find the work done by the force of gravity:
P2
LP1
S
S
w ~ dl =
L0
u0
1-w sin u2R du = - wR
L0
u0
sin u du
= -wR11 - cos u02
The workS done by gravity is indeed the negative of the work done
by force F that we calculated in part (c). Gravity does negative work
because the force pulls downward while Throcky moves upward.
As we saw earlier, R11 - cos u02 is equal to h, the increase
in Throcky’s height during the displacement. So the work done
by gravity along the curved path is - mgh, the same work that
gravity would have done if Throcky had moved straight upward
a distance h. This is an example of a more general result that we’ll
prove in Section 7.1.
teSt your underStanding of Section 6.3 In Example 5.20 (Section 5.4)
we examined a conical pendulum. The speed of the pendulum bob remains constant
as it travels around the circle shown in Fig. 5.32a. (a) Over one complete circle, how
much work does the tension force F do on the bob? (i) A positive amount; (ii) a negative
amount; (iii) zero. (b) Over one complete circle, how much work does the weight do on
the bob? (i) A positive amount; (ii) a negative amount; (iii) zero. ❙
6.4 power
The definition of work makes no reference to the passage of time. If you lift a
barbell weighing 100 N through a vertical distance of 1.0 m at constant velocity,
you do 1100 N211.0 m2 = 100 J of work whether it takes you 1 second, 1 hour,
or 1 year to do it. But often we need to know how quickly work is done. We describe this in terms of power. In ordinary conversation the word “power” is often
synonymous with “energy” or “force.” In physics we use a much more precise
definition: Power is the time rate at which work is done. Like work and energy,
power is a scalar quantity.
The average work done per unit time, or average power Pav, is defined to be
Average power during
time interval ∆t
Pav =
∆W
∆t
Work done during time interval
Duration of time interval
∆W
dW
=
∆t S 0 ∆t
dt
P = lim
Time rate of
doing work
t = 5s
Work you do on the box
to lift it in 5 s:
W = 100 J
Your power output:
100 J
W
P =
=
= 20 W
t
5s
(6.15)
t = 0
The rate at which work is done might not be constant. We define instantaneous
power P as the quotient in Eq. (6.15) as ∆t approaches zero:
Instantaneous
power
6.25 The same amount of work is done in
both of these situations, but the power (the
rate at which work is done) is different.
t = 1s
Work you do on the same
box to lift it the same
distance in 1 s:
W = 100 J
Your power output:
100 J
W
P =
=
= 100 W
t
1s
(6.16)
Average power over infinitesimally short time interval
The SI unit of power is the watt (W), named for the English inventor James
Watt. One watt equals 1 joule per second: 1 W = 1 J>s (Fig. 6.25). The kilowatt
11 kW = 103 W2 and the megawatt 11 MW = 106 W2 are also commonly used.
t = 0
190
ChaPtEr 6 Work and Kinetic Energy
Another common unit of power is the horsepower (hp) (Fig. 6.26). The value
of this unit derives from experiments by James Watt, who measured that in one
minute a horse could do an amount of work equivalent to lifting 33,000 pounds
(lb) a distance of 1 foot (ft), or 33,000 ft # lb. Thus 1 hp = 33,000 ft # lb>min.
Using 1 ft = 0.3048 m, 1 lb = 4.448 N, and 1 min = 60 s, we can show that
6.26 A one-horsepower (746-W)
propulsion system.
1 hp = 746 W = 0.746 kW
BIO application Muscle Power Skeletal
muscles provide the power that makes animals
move. Muscle fibers that rely on anaerobic
metabolism do not require oxygen; they produce large amounts of power but are useful for
short sprints only. Muscle fibers that metabolize aerobically use oxygen and produce
smaller amounts of power for long intervals.
Both fiber types are visible in a fish fillet: The
pale (anaerobic) muscle is used for brief bursts
of speed, while the darker (aerobic) muscle is
used for sustained swimming.
Anaerobic muscle
The watt is a familiar unit of electrical power; a 100-W light bulb converts
100 J of electrical energy into light and heat each second. But there’s nothing
inherently electrical about a watt. A light bulb could be rated in horsepower, and
an engine can be rated in kilowatts.
The kilowatt-hour 1kW # h2 is the usual commercial unit of electrical energy.
One kilowatt-hour is the total work done in 1 hour (3600 s) when the power is
1 kilowatt 1103 J>s2, so
1 kW # h = 1103 J>s213600 s2 = 3.6 * 106 J = 3.6 MJ
The kilowatt-hour is a unit of work or energy, not power.
In mechanics we
can also express power in terms of force and velocity. SupS
S
pose that a force F acts onSa body while it undergoes a vector displacement ∆s .
S
If FŒ is the component of F tangent to the path (parallel to ∆s ), then the work
done by the force is ∆W = FŒ ∆s. The average power is
Pav =
FŒ ∆s
∆s
= FŒ
= FΠvav
∆t
∆t
(6.17)
Instantaneous power P is the limit of this expression as ∆t S 0:
P = FΠv
(6.18)
where v is the magnitude of the instantaneous velocity. We can also express
Eq. (6.18) in terms of the scalar product:
Instantaneous power
for a force doing work
on a particle
S
S
P = F~v
Force that acts on particle
Velocity of particle
(6.19)
Aerobic muscle
Solution
ExamplE 6.9 ForcE and powEr
Each of the four jet engines on an Airbus A380 airliner develops
a thrust (a forward force on the airliner) of 322,000 N (72,000 lb).
When the airplane is flying at 250 m>s (900 km>h, or roughly
560 mi>h), what horsepower does each engine develop?
6.27 (a) Propeller-driven and (b) jet airliners.
(a)
(b)
solution
idEntifY, sEt uP, and ExEcutE: Our target variable is the instantaneous power P, which is the rate at which the thrust does work.
We use Eq. (6.18). The thrust is in the direction of motion, so FΠis
just equal to the thrust. At v = 250 m>s, the power developed by
each engine is
5
7
P = FŒv = 13.22 * 10 N21250 m>s2 = 8.05 * 10 W
= 18.05 * 107 W2
1 hp
= 108,000 hp
746 W
EvaluatE: The speed of modern airliners is directly related to the
power of their engines (Fig. 6.27). The largest propeller-driven
airliners of the 1950s had engines that each developed about 3400 hp
12.5 * 106 W2, giving them maximum speeds of about 600 km>h
(370 mi>h). Each engine on an Airbus A380 develops more than
30 times more power, enabling it to fly at about 900 km>h (560 mi>h)
and to carry a much heavier load.
If the engines are at maximum thrust while the airliner is at
rest on the ground so that v = 0, the engines develop zero power.
Force and power are not the same thing!
6.4 power
Solution
ExamplE 6.10
a “poWEr climb”
A 50.0-kg marathon runner runs up the stairs to the top of Chicago’s
443-m-tall Willis Tower, the second tallest building in the United
States (Fig. 6.28). To lift herself to the top in 15.0 minutes, what
must be her average power output? Express your answer in watts,
in kilowatts, and in horsepower.
Solution
iDentify and Set uP: We’ll treat the runner as a particle of
mass m. Her average power output Pav must be enough to lift her
at constant speed against gravity.
We can find Pav in two ways: (1) by determining how much
work she must do and dividing that quantity by the elapsed time,
6.28 How much power is required to run up the stairs of
Chicago’s Willis Tower in 15 minutes?
191
as in Eq. (6.15), or (2) by calculating the average upward force
she must exert (in the direction of the climb) and multiplying that
quantity by her upward velocity, as in Eq. (6.17).
exeCute: (1) As in Example 6.8, lifting a mass m against gravity
requires an amount of work equal to the weight mg multiplied by
the height h it is lifted. Hence the work the runner must do is
W = mgh = 150.0 kg219.80 m>s221443 m2
= 2.17 * 105 J
She does this work in a time 15.0 min = 900 s, so from Eq. (6.15)
the average power is
Pav =
2.17 * 105 J
= 241 W = 0.241 kW = 0.323 hp
900 s
(2) The force exerted is vertical and the average vertical
component of velocity is 1443 m2>1900 s2 = 0.492 m>s, so from
Eq. (6.17) the average power is
Pav = FŒvav = 1mg2vav
= 150.0 kg219.80 m>s2210.492 m>s2 = 241 W
which is the same result as before.
evaluate: The runner’s total power output will be several times
greater than 241 W. The reason is that the runner isn’t really a particle but a collection of parts that exert forces on each other and do
work, such as the work done to inhale and exhale and to make her
arms and legs swing. What we’ve calculated is only the part of her
power output that lifts her to the top of the building.
teSt your unDerStanDing of SeCtion 6.4 The air surrounding an
airplane in flight exerts a drag force that acts opposite to the airplane’s motion. When
the Airbus A380 in Example 6.9 is flying in a straight line at a constant altitude at a
constant 250 m>s, what is the rate at which the drag force does work on it? (i) 432,000 hp;
(ii) 108,000 hp; (iii) 0; (iv) - 108,000 hp; (v) -432,000 hp. ❙
Chapter
6 Summary
S
Work done by a force: When a constant force F
acts on a particle that undergoes a straight-line
S
displacement s , the work done by the force on the
particle
is defined to be the scalar product
S
S
of F and s . The unit of work in SI units is
1 joule = 1 newton@meter 11 J = 1 N # m2. Work
is a scalar quantity; it can be positive or negative,
but it has no direction in space. (See Examples 6.1
and 6.2.)
Kinetic energy: The kinetic energy K of a particle
equals the amount of work required to accelerate
the particle from rest to speed v. It is also equal
to the amount of work the particle can do in the
process of being brought to rest. Kinetic energy is
a scalar that has no direction in space; it is always
positive or zero. Its units are the same as the units
of work: 1 J = 1 N # m = 1 kg # m2>s2.
The work–energy theorem: When forces act on a
particle while it undergoes a displacement, the particle’s kinetic energy changes by an amount equal to
the total work done on the particle by all the forces.
This relationship, called the work–energy theorem,
is valid whether the forces are constant or varying
and whether the particle moves along a straight or
curved path. It is applicable only to bodies that can
be treated as particles. (See Examples 6.3–6.5.)
Work done by a varying force or on a curved path:
When a force varies during a straight-line displacement, the work done by the force is given by
an integral, Eq. (6.7). (See Examples 6.6 and 6.7.)
When a particle follows
a curved path, the work
S
done on it by a force F is given by an integral that
involves the angle f between the force and the
displacement. This expression is valid even if the
force magnitude and the angle f vary during the
displacement. (See Example 6.8.)
Power: Power is the time rate of doing work. The
average power Pav is the amount of work ∆W done
in time ∆t divided by that time. The instantaneous
power is the limit of the
average power as ∆t goes
S
to zero. When a force F acts on a particle moving
S
with velocity v, the instantaneous power (the rate
at which
the force does work) is the scalar product
S
S
of F and v. Like work and kinetic energy, power
is a scalar quantity. The SI unit of power is
1 watt = 1 joule>second 11 W = 1 J>s2.
(See Examples 6.9 and 6.10.)
192
SolutionS to all exampleS
S
S
S
W = F ~ s = Fs cos f
S
S
f = angle between F and s
(6.2), (6.3)
F
F#
W = FŒs
= (F cos f)s
f
FΠ= F cos f
K = 12 mv 2
m
(6.5)
2m
S
v
S
v
Doubling m doubles K.
m
m
S
v
S
2v
Doubling v quadruples K.
Wtot = K2 - K1 = ∆K
(6.6)
m
v1
Wtot = Total work done on
particle along path
m
K1 = 12 mv12
K2 =
1
mv22
2
= K1 + Wtot
x2
W =
Lx1
Fx dx
P2
W =
LP1
S
(6.7)
S
F ~ dl
P2
=
Pav =
LP1
P2
F cos f dl =
∆W
∆t
P = lim
∆t S 0
S
Area = Work done by
force during displacement
Fx
S
P = F~v
∆W
dW
=
∆t
dt
LP1
(6.14)
O
(6.15)
t = 5s
x1
x2
FΠdl
(6.16)
(6.19)
t = 0
Work you do on the
box to lift it in 5 s:
W = 100 J
Your power output:
100 J
W
P =
=
t
5s
= 20 W
v2
x
Discussion Questions
Solution
bridging problEm
a spring that disobEys hooKE’s laW
Consider a hanging spring of negligible mass that does not obey
Hooke’s law. When the spring is pulled downward by a distance x,
the spring exerts an upward force of magnitude ax 2, where a is a positive constant. Initially the hanging spring is relaxed (not extended).
We then attach a block of mass m to the spring and release the block.
The block stretches the spring as it falls (Fig. 6.29). (a) How fast
is the block moving when it has fallen a distance x1? (b) At what
rate does the spring do work on the block at this point? (c) Find the
maximum distance x2 that the spring stretches. (d) Will the block
remain at the point found in part (c)?
Solution guiDe
iDentify and Set uP
1. The spring force in this problem isn’t constant, so you have to
use the work–energy theorem. You’ll also need Eq. (6.7) to
find the work done by the spring over a given displacement.
6.29 The block is attached to a spring that does not obey
Hooke’s law.
Block is attached
to end of relaxed
spring, then released.
m
x
Positive x is
downward.
193
2. Draw a free-body diagram for the block, including your choice
of coordinate axes. Note that x represents how far the spring is
stretched, so choose the positive x-direction to be downward,
as in Fig. 6.29. On your coordinate axis, label the points x = x1
and x = x2.
3. Make a list of the unknown quantities, and decide which of
these are the target variables.
exeCute
4. Calculate the work done on the block by the spring as the block
falls an arbitrary distance x. (The integral isn’t a difficult one.
Use Appendix B if you need a reminder.) Is the work done by
the spring positive, negative, or zero?
5. Calculate the work done on the block by any other forces as the
block falls an arbitrary distance x. Is this work positive, negative, or zero?
6. Use the work–energy theorem to find the target variables.
(You’ll also need an equation for power.) Hint: When the spring
is at its maximum stretch, what is the speed of the block?
7. To answer part (d), consider the net force that acts on the block
when it is at the point found in part (c).
evaluate
8. We learned in Section 2.5 that after an object dropped from
rest has fallen freely a distance x1, its speed is 12gx1. Use this
to decide whether your answer in part (a) makes sense. In addition, ask yourself whether the algebraic sign of your answer in
part (b) makes sense.
9. Find the value of x where the net force on the block would be
zero. How does this compare to your result for x2? Is this consistent with your answer in part (d)?
Problems
For assigned homework and other learning materials, go to MasteringPhysics®.
., .., ...: Difficulty levels. CP: Cumulative problems incorporating material from earlier chapters. CALC: Problems requiring calculus.
DATA: Problems involving real data, scientific evidence, experimental design, and/or statistical reasoning. bio: Biosciences problems.
DiSCuSSion QueStionS
Q6.1 The sign of many physical quantities depends on the choice
of coordinates. For example, ay for free-fall motion can be negative or positive, depending on whether we choose upward or
downward as positive. Is the same true of work? In other words,
can we make positive work negative by a different choice of coordinates? Explain.
Q6.2 An elevator is hoisted by its cables at constant speed. Is the
total work done on the elevator positive, negative, or zero? Explain.
Q6.3 A rope tied to a body is pulled, causing the body to accelerate. But according to Newton’s third law, the body pulls back on
the rope with a force of equal magnitude and opposite direction.
Is the total work done then zero? If so, how can the body’s kinetic
energy change? Explain.
Q6.4 If it takes total work W to give an object a speed v and kinetic energy K, starting from rest, what will be the object’s speed
(in terms of v) and kinetic energy (in terms of K) if we do twice as
much work on it, again starting from rest?
Q6.5 If there is a net nonzero force on a moving object, can the
total work done on the object be zero? Explain, using an example.
Q6.6 In Example 5.5 (Section 5.1), how does the work done on
the bucket by the tension in the cable compare with the work done
on the cart by the tension in the cable?
Q6.7 In the conical pendulum of Example 5.20 (Section 5.4),
which of the forces do work on the bob while it is swinging?
Q6.8 For the cases shown in Fig. Q6.8, the object is released from
rest at the top and feels no friction or air resistance. In which (if any)
cases will the mass have (i) the greatest speed at the bottom and
(ii) the most work done on it by the time it reaches the bottom?
Figure Q6.8
(a)
(b)
m
h
(c)
m
h
2m
h
194
Chapter 6 Work and Kinetic energy
S
Q6.9 A force F is in the x-direction and has a magnitude that
depends on x. Sketch a possible graph of F versus x such that the
force does zero work on an object that moves from x1 to x2, even
though the force magnitude is not zero at all x in this range.
Q6.10 Does a car’s kinetic energy change more when the car
speeds up from 10 to 15 m>s or from 15 to 20 m>s? Explain.
Q6.11 A falling brick has a mass of 1.5 kg and is moving straight
downward with a speed of 5.0 m>s. A 1.5-kg physics book is sliding across the floor with a speed of 5.0 m>s. A 1.5-kg melon is
traveling with a horizontal velocity component 3.0 m>s to the right
and a vertical component 4.0 m>s upward. Do all of these objects
have the same velocity? Do all of them have the same kinetic energy? For both questions, give your reasoning.
Q6.12 Can the total work done on an object during a displacement be negative? Explain. If the total work is negative, can its
magnitude be larger than the initial kinetic energy of the object?
Explain.
Q6.13 A net force acts on an object and accelerates it from rest
to a speed v1. In doing so, the force does an amount of work W1.
By what factor must the work done on the object be increased to
produce three times the final speed, with the object again starting
from rest?
Q6.14 A truck speeding down the highway has a lot of kinetic
energy relative to a stopped state trooper but no kinetic energy
relative to the truck driver. In these two frames of reference, is the
same amount of work required to stop the truck? Explain.
Q6.15 You are holding a briefcase by the handle, with your arm
straight down by your side. Does the force your hand exerts do
work on the briefcase when (a) you walk at a constant speed down
a horizontal hallway and (b) you ride an escalator from the first
to second floor of a building? In both cases justify your answer.
Q6.16 When a book slides along a tabletop, the force of friction does negative work on it. Can friction ever do positive work?
Explain. (Hint: Think of a box in the back of an accelerating
truck.)
Q6.17 Time yourself while running up a flight of steps, and
compute the average rate at which you do work against the force
of gravity. Express your answer in watts and in horsepower.
Q6.18 Fractured Physics. Many terms from physics are badly
misused in everyday language. In both cases, explain the errors
involved. (a) A strong person is called powerful. What is wrong
with this use of power? (b) When a worker carries a bag of concrete along a level construction site, people say he did a lot of
work. Did he?
Q6.19 An advertisement for a portable electrical generating unit
claims that the unit’s diesel engine produces 28,000 hp to drive an
electrical generator that produces 30 MW of electrical power. Is
this possible? Explain.
Q6.20 A car speeds up while the engine delivers constant power.
Is the acceleration greater at the beginning of this process or at the
end? Explain.
Q6.21 Consider a graph of instantaneous power versus time,
with the vertical P-axis starting at P = 0. What is the physical
significance of the area under the P-versus-t curve between vertical lines at t1 and t2 ? How could you find the average power from
the graph? Draw a P-versus-t curve that consists of two straightline sections and for which the peak power is equal to twice the
average power.
Q6.22 A nonzero net force acts on an object. Is it possible for any
of the following quantities to be constant: the object’s (a) speed;
(b) velocity; (c) kinetic energy?
Q6.23 When a certain force is applied to an ideal spring, the
spring stretches a distance x from its unstretched length and does
work W. If instead twice the force is applied, what distance (in terms
of x) does the spring stretch from its unstretched length, and how
much work (in terms of W) is required to stretch it this distance?
Q6.24 If work W is required to stretch a spring a distance x from
its unstretched length, what work (in terms of W) is required to
stretch the spring an additional distance x?
exerCises
Section 6.1 Work
6.1
.
You push your physics book 1.50 m along a horizontal tabletop with a horizontal push of 2.40 N while the opposing force of
friction is 0.600 N. How much work does each of the following
forces do on the book: (a) your 2.40-N push, (b) the friction force,
(c) the normal force from the tabletop, and (d) gravity? (e) What is
the net work done on the book?
6.2 . Using a cable with a tension of 1350 N, a tow truck pulls a
car 5.00 km along a horizontal roadway. (a) How much work does
the cable do on the car if it pulls horizontally? If it pulls at 35.0°
above the horizontal? (b) How much work does the cable do on the
tow truck in both cases of part (a)? (c) How much work does gravity do on the car in part (a)?
6.3 . A factory worker pushes a 30.0-kg crate a distance of 4.5 m
along a level floor at constant velocity by pushing horizontally on
it. The coefficient of kinetic friction between the crate and the
floor is 0.25. (a) What magnitude of force must the worker apply?
(b) How much work is done on the crate by this force? (c) How
much work is done on the crate by friction? (d) How much work is
done on the crate by the normal force? By gravity? (e) What is the
total work done on the crate?
6.4 .. Suppose the worker in Exercise 6.3 pushes downward at
an angle of 30° below the horizontal. (a) What magnitude of force
must the worker apply to move the crate at constant velocity?
(b) How much work is done on the crate by this force when the
crate is pushed a distance of 4.5 m? (c) How much work is done on
the crate by friction during this displacement? (d) How much work
is done on the crate by the normal force? By gravity? (e) What is
the total work done on the crate?
6.5 .. A 75.0-kg painter climbs a ladder that is 2.75 m long and
leans against a vertical wall. The ladder makes a 30.0° angle with
the wall. (a) How much work does gravity do on the painter?
(b) Does the answer to part (a) depend on whether the painter
climbs at constant speed or accelerates up the ladder?
6.6 .. Two tugboats pull a disabled supertanker. Each tug exerts
a constant force of 1.80 * 106 N, one 14° west of north and the
other 14° east of north, as they pull the tanker 0.75 km toward the
north. What is the total work they do on the supertanker?
6.7 . Two blocks are con- Figure E6.7
nected by a very light string
passing over a massless and
20.0
N
frictionless pulley (Fig. E6.7).
Traveling at constant speed,
the 20.0-N block moves 75.0 cm
12.0
to the right and the 12.0-N
N
block moves 75.0 cm downward. How much work is done
(a) on the 12.0-N block by (i) gravity and (ii) the tension in the
string? (b) How much work is done on the 20.0-N block by
exercises
(i) gravity, (ii) the tension in the string, (iii) friction, and (iv) the
normal force? (c) Find the total work done on each block.
6.8 .. A loaded grocery cart is rolling
across a parking lot in a strong
S
wind. You apply a constant force F = 130 N2dn − 140 N2en to the
S
cart as it undergoes a displacement s = 1-9.0 m2dn − 13.0 m2en.
How much work does the force you apply do on the grocery cart?
6.9 . A 0.800-kg ball is tied to the end of a string 1.60 m long
and swung in a vertical circle. (a) During one complete circle,
starting anywhere, calculate the total work done on the ball by
(i) the tension in the string and (ii) gravity. (b) Repeat part (a) for
motion along the semicircle from the lowest to the highest point
on the path.
6.10 .. A 12.0-kg package in a mail-sorting room slides 2.00 m
down a chute that is inclined at 53.0° below the horizontal. The
coefficient of kinetic friction between the package and the chute’s
surface is 0.40. Calculate the work done on the package by
(a) friction, (b) gravity, and (c) the normal force. (d) What is the
net work done on the package?
6.11 . A 128.0-N carton is pulled up a frictionless baggage ramp
inclined at 30.0° above the horizontal by a rope exerting a 72.0-N
pull parallel to the ramp’s surface. If the carton travels 5.20 m
along the surface of the ramp, calculate the work done on it by
(a) the rope, (b) gravity, and (c) the normal force of the ramp.
(d) What is the net work done on the carton? (e) Suppose that the
rope is angled at 50.0° above the horizontal, instead of being parallel to the ramp’s surface. How much work does the rope do on
the carton in this case?
6.12 .. A boxed 10.0-kg computer monitor is dragged by friction 5.50 m upward along a conveyor belt inclined at an angle of
36.9° above the horizontal. If the monitor’s speed is a constant
2.10 cm>s, how much work is done on the monitor by (a) friction,
(b) gravity, and (c) the normal force of the conveyor belt?
6.13 .. A large crate sits on the floor of a warehouse. Paul and
Bob apply constant horizontal forces to the crate. The force applied
by Paul has magnitude 48.0 N and direction 61.0° south of west.
How much work does Paul’s force do during a displacement of the
crate that is 12.0 m in the direction 22.0°
east of north?
S
6.14 .. You apply a constant force F = 1- 68.0 N2dn + 136.0 N2en
to a 380-kg car as the car travels 48.0 m in a direction that is
240.0° counterclockwise from the + x-axis. How much work does
the force you apply do on the car?
6.15 .. On a farm, you are pushing on a stubborn pig with a constant horizontal force with magnitude 30.0 N and direction 37.0°
counterclockwise from the + x-axis. How much work does this
S
force do during a displacement of the pig that is (a) s = 15.00 m2dn;
S
S
(b) s = −16.00 m2en; (c) s = −12.00 m2dn + 14.00 m2en?
Section 6.2 Kinetic Energy and the Work–Energy
Theorem
6.16 .. A 1.50-kg book is sliding along a rough horizontal surface. At point A it is moving at 3.21 m>s, and at point B it has
slowed to 1.25 m>s. (a) How much work was done on the book between A and B? (b) If - 0.750 J of work is done on the book from
B to C, how fast is it moving at point C? (c) How fast would it be
moving at C if + 0.750 J of work was done on it from B to C?
6.17 .. bio Animal Energy. Adult cheetahs, the fastest of the
great cats, have a mass of about 70 kg and have been clocked to
run at up to 72 mi>h 132 m>s2. (a) How many joules of kinetic
energy does such a swift cheetah have? (b) By what factor would
its kinetic energy change if its speed were doubled?
195
. Some Typical Kinetic Energies. (a) In the Bohr model
of the atom, the ground-state electron in hydrogen has an orbital
speed of 2190 km>s. What is its kinetic energy? (Consult
Appendix F.) (b) If you drop a 1.0-kg weight (about 2 lb) from a
height of 1.0 m, how many joules of kinetic energy will it have
when it reaches the ground? (c) Is it reasonable that a 30-kg child
could run fast enough to have 100 J of kinetic energy?
6.19 . Meteor Crater. About 50,000 years ago, a meteor
crashed into the earth near present-day Flagstaff, Arizona.
Measurements from 2005 estimate that this meteor had a mass
of about 1.4 * 108 kg (around 150,000 tons) and hit the ground
at a speed of 12 km>s. (a) How much kinetic energy did this
meteor deliver to the ground? (b) How does this energy compare to the energy released by a 1.0-megaton nuclear bomb?
(A megaton bomb releases the same amount of energy as a million tons of TNT, and 1.0 ton of TNT releases 4.184 * 109 J of
energy.)
6.20 . A 4.80-kg watermelon is dropped from rest from the roof
of an 18.0-m-tall building and feels no appreciable air resistance.
(a) Calculate the work done by gravity on the watermelon during
its displacement from the roof to the ground. (b) Just before it
strikes the ground, what is the watermelon’s (i) kinetic energy and
(ii) speed? (c) Which of the answers in parts (a) and (b) would be
different if there were appreciable air resistance?
6.21 .. Use the work–energy theorem to solve each of these
problems. You can use Newton’s laws to check your answers.
Neglect air resistance in all cases. (a) A branch falls from the top
of a 95.0-m-tall redwood tree, starting from rest. How fast is it
moving when it reaches the ground? (b) A volcano ejects a boulder
directly upward 525 m into the air. How fast was the boulder moving just as it left the volcano?
6.22 .. Use the work–energy theorem to solve each of these
problems. You can use Newton’s laws to check your answers. (a) A
skier moving at 5.00 m>s encounters a long, rough horizontal
patch of snow having a coefficient of kinetic friction of 0.220 with
her skis. How far does she travel on this patch before stopping?
(b) Suppose the rough patch in part (a) was only 2.90 m long. How
fast would the skier be moving when she reached the end of the
patch? (c) At the base of a frictionless icy hill that rises at 25.0°
above the horizontal, a toboggan has a speed of 12.0 m>s toward
the hill. How high vertically above the base will it go before
stopping?
6.23 .. You are a member of an Alpine Rescue Team. You must
project a box of supplies up an incline of constant slope angle a so
that it reaches a stranded skier who is a vertical distance h above
the bottom of the incline. The incline is slippery, but there is some
friction present, with kinetic friction coefficient mk. Use the work–
energy theorem to calculate the minimum speed you must give the
box at the bottom of the incline so that it will reach the skier.
Express your answer in terms of g, h, mk, and a.
6.24 .. You throw a 3.00-N rock vertically into the air from
ground level. You observe that when it is 15.0 m above the ground,
it is traveling at 25.0 m>s upward. Use the work–energy theorem
to find (a) the rock’s speed just as it left the ground and (b) its
maximum height.
6.25 . A sled with mass 12.00 kg moves in a straight line on a
frictionless, horizontal surface. At one point in its path, its speed
is 4.00 m>s; after it has traveled 2.50 m beyond this point, its
speed is 6.00 m>s. Use the work–energy theorem to find the force
acting on the sled, assuming that this force is constant and that it
acts in the direction of the sled’s motion.
6.18
196
Chapter 6 Work and Kinetic energy
6.26 .. A mass m slides down a smooth inclined plane from an
initial vertical height h, making an angle a with the horizontal.
(a) The work done by a force is the sum of the work done by the
components of the force. Consider the components of gravity parallel and perpendicular to the surface of the plane. Calculate the work
done on the mass by each of the components, and use these results
to show that the work done by gravity is exactly the same as if the
mass had fallen straight down through the air from a height h. (b) Use
the work–energy theorem to prove that the speed of the mass at the
bottom of the incline is the same as if the mass had been dropped
from height h, independent of the angle a of the incline. Explain
how this speed can be independent of the slope angle. (c) Use the
results of part (b) to find the speed of a rock that slides down an icy
frictionless hill, starting from rest 15.0 m above the bottom.
6.27 . A 12-pack of Omni-Cola (mass 4.30 kg) is initially at rest
on a horizontal floor. It is then pushed in a straight line for 1.20 m
by a trained dog that exerts a horizontal force with magnitude
36.0 N. Use the work–energy theorem to find the final speed of
the 12-pack if (a) there is no friction between the 12-pack and
the floor, and (b) the coefficient of kinetic friction between the
12-pack and the floor is 0.30.
6.28 .. A soccer ball with mass 0.420 kg is initially moving with
speed 2.00 m>s. A soccer player kicks the ball, exerting a constant
force of magnitude 40.0 N in the same direction as the ball’s
motion. Over what distance must the player’s foot be in contact
with the ball to increase the ball’s speed to 6.00 m>s ?
6.29 . A little red wagon with mass 7.00 kg moves in a straight
line on a frictionless horizontal surface. It has an initial speed of
4.00 m>s and then is pushed 3.0 m in the direction of the initial
velocity by a force with a magnitude of 10.0 N. (a) Use the work–
energy theorem to calculate the wagon’s final speed. (b) Calculate
the acceleration produced by the force. Use this acceleration in the
kinematic relationships of Chapter 2 to calculate the wagon’s final
speed. Compare this result to that calculated in part (a).
6.30 .. A block of ice with mass 2.00 kg slides 1.35 m down an
inclined plane that slopes downward at an angle of 36.9° below the
horizontal. If the block of ice starts from rest, what is its final
speed? Ignore friction.
6.31 . Stopping Distance. A car is traveling on a level road
with speed v0 at the instant when the brakes lock, so that the tires
slide rather than roll. (a) Use the work–energy theorem to calculate the minimum stopping distance of the car in terms of v0, g,
and the coefficient of kinetic friction mk between the tires and the
road. (b) By what factor would the minimum stopping distance
change if (i) the coefficient of kinetic friction were doubled, or
(ii) the initial speed were doubled, or (iii) both the coefficient of
kinetic friction and the initial speed were doubled?
6.32 .. A 30.0-kg crate is initially moving with a velocity that
has magnitude 3.90 m>s in a direction 37.0o west of north. How
much work must be done on the crate to change its velocity to
5.62 m>s in a direction 63.0° south of east?
Section 6.3 Work and Energy with Varying Forces
. bio Heart Repair. A surgeon is using material from a
donated heart to repair a patient’s damaged aorta and needs to know
the elastic characteristics of this aortal material. Tests performed on
a 16.0-cm strip of the donated aorta reveal that it stretches 3.75 cm
when a 1.50-N pull is exerted on it. (a) What is the force constant of
this strip of aortal material? (b) If the maximum distance it will be
able to stretch when it replaces the aorta in the damaged heart is
1.14 cm, what is the greatest force it will be able to exert there?
6.33
6.34 .. To stretch a spring 3.00 cm from its unstretched length,
12.0 J of work must be done. (a) What is the force constant of this
spring? (b) What magnitude force is needed to stretch the spring
3.00 cm from its unstretched length? (c) How much work must be
done to compress this spring 4.00 cm from its unstretched length,
and what force is needed to compress it this distance?
6.35 . Three identical 8.50-kg masses are hung Figure E6.35
by three identical springs (Fig. E6.35). Each
spring has a force constant of 7.80 kN>m and was
12.0 cm long before any masses were attached
to it. (a) Draw a free-body diagram of each mass.
(b) How long is each spring when hanging as
shown? (Hint: First isolate only the bottom mass.
Then treat the bottom two masses as a system.
Finally, treat all three masses as a system.)
6.36
.
S
A child applies a force F Figure E6.36
parallel to the x-axis to a 10.0Fx (N)
kg sled moving on the frozen
10
surface of a small pond. As the
child controls the speed of
the sled, the x-component of the
5
force she applies varies with
the x-coordinate of the sled as
shown in Fig. E6.36.
Calculate
x (m)
S
0
4
8
12
the work done by F when the
sled moves (a) from x = 0 to
x = 8.0 m; (b) from x = 8.0 m to x = 12.0 m; (c) from x = 0 to
12.0 m.
6.37 .. Suppose the sled in Exercise 6.36 is initially at rest at
x = 0. Use the work–energy theorem to find the speed of the sled
at (a) x = 8.0 m and (b) x = 12.0 m. Ignore friction between the
sled and the surface of the pond.
6.38 .. A spring of force constant 300.0 N>m and unstretched
length 0.240 m is stretched by two forces, pulling in opposite
directions at opposite ends of the spring, that increase to 15.0 N.
How long will the spring now be, and how much work was
required to stretch it that distance?
6.39 .. A 6.0-kg box moving at 3.0 m>s on a horizontal, frictionless surface runs into a light spring of force constant 75 N>cm.
Use the work–energy theorem to find the maximum compression
of the spring.
6.40 .. Leg Presses. As part of your daily workout, you lie on
your back and push with your feet against a platform attached to
two stiff springs arranged side by side so that they are parallel
to each other. When you push the platform, you compress the
springs. You do 80.0 J of work when you compress the springs
0.200 m from their uncompressed length. (a) What magnitude of
force must you apply to hold the platform in this position? (b) How
much additional work must you do to move the platform 0.200 m
farther, and what maximum force must you apply?
6.41 .. (a) In Example 6.7 (Section 6.3) it was calculated that
with the air track turned off, the glider travels 8.6 cm before it
stops instantaneously. How large would the coefficient of static
friction ms have to be to keep the glider from springing back to the
left? (b) If the coefficient of static friction between the glider and
the track is ms = 0.60, what is the maximum initial speed v1 that
the glider can be given and still remain at rest after it stops instantaneously? With the air track turned off, the coefficient of kinetic
friction is mk = 0.47.
exercises
. A 4.00-kg block of ice is placed against a horizontal
spring that has force constant k = 200 N>m and is compressed
0.025 m. The spring is released and accelerates the block along
a horizontal surface. Ignore friction and the mass of the spring.
(a) Calculate the work done on the block by the spring during the
motion of the block from its initial position to where the spring
has returned to its uncompressed length. (b) What is the speed of
the block after it leaves
the spring?
S
6.43 . A force F is applied to Figure E6.43
a 2.0-kg, radio-controlled model
Fx (N)
car parallel to the x-axis as it
2
moves along a straight track.
1
6
The x-component of the force
x (m)
0
varies with the x-coordinate of
7
-1 1 2 3 4 5
the car (Fig. E6.43). Calculate
S
-2
the work done by the force F
when the car moves from
(a) x = 0 to x = 3.0 m; (b) x = 3.0 m to x = 4.0 m; (c) x =
x = 4.0 m to x = 7.0 m; (d) x = 0 to x = 7.0 m; (e) x = 7.0 m to
x = 2.0 m.
6.44 . Suppose the S2.0-kg model car in Exercise 6.43 is initially
at rest at x = 0 and F is the net force acting on it. Use the work–
energy theorem to find the speed of the car at (a) x = 3.0 m;
(b) x = 4.0 m; (c) x = 7.0 m.
6.45 .. At a waterpark, sleds with riders are sent along a slippery, horizontal surface by the release of a large compressed
spring. The spring, with force constant k = 40.0 N>cm and negligible mass, rests on the frictionless horizontal surface. One end is
in contact with a stationary wall. A sled and rider with total mass
70.0 kg are pushed against the other end, compressing the spring
0.375 m. The sled is then released with zero initial velocity. What
is the sled’s speed when the spring (a) returns to its uncompressed
length and (b) is still compressed 0.200 m?
6.46 . Half of a Spring. (a) Suppose you cut a massless ideal
spring in half. If the full spring had a force constant k, what is the
force constant of each half, in terms of k? (Hint: Think of the original spring as two equal halves, each producing the same force as
the entire spring. Do you see why the forces must be equal?) (b) If
you cut the spring into three equal segments instead, what is the
force constant of each one, in terms of k?
6.47 .. A small glider is placed against a compressed spring at
the bottom of an air track that slopes upward at an angle of 40.0°
above the horizontal. The glider has mass 0.0900 kg. The spring
has k = 640 N>m and negligible mass. When the spring is released, the glider travels a maximum distance of 1.80 m along the
air track before sliding back down. Before reaching this maximum
distance, the glider loses contact with the spring. (a) What distance was the spring originally compressed? (b) When the glider
has traveled along the air track 0.80 m from its initial position
against the compressed spring, is it still in contact with the spring?
What is the kinetic energy of the glider at this point?
6.48 .. An ingenious bricklayer builds a device for shooting
bricks up to the top of the wall where he is working. He places a
brick on a vertical compressed spring with force constant
k = 450 N>m and negligible mass. When the spring is released,
the brick is propelled upward. If the brick has mass 1.80 kg and
is to reach a maximum height of 3.6 m above its initial position
on the compressed spring, what distance must the bricklayer
compress the spring initially? (The brick loses contact with the
spring when the spring returns to its uncompressed length.
Why?)
6.42
197
.. CALC A force in the + x-direction with magnitude
F1x2 = 18.0 N - 10.530 N>m2x is applied to a 6.00-kg box that
is sitting on the horizontal, frictionless surface of a frozen lake.
F1x2 is the only horizontal force on the box. If the box is initially
at rest at x = 0, what is its speed after it has traveled 14.0 m?
6.49
Section 6.4 Power
6.50 .. A crate on a motorized cart starts from rest and moves
with a constant eastward acceleration of a = 2.80 m>s2. A worker
assists the cart by pushing on the crate with a force that is eastward and has magnitude that depends on time according to
F1t2 = 15.40 N>s2t. What is the instantaneous power supplied by
this force at t = 5.00 s?
6.51 . How many joules of energy does a 100-watt light bulb use
per hour? How fast would a 70-kg person have to run to have that
amount of kinetic energy?
6.52 .. BIO Should You Walk or Run? It is 5.0 km from your
home to the physics lab. As part of your physical fitness program,
you could run that distance at 10 km>h (which uses up energy at
the rate of 700 W), or you could walk it leisurely at 3.0 km>h
(which uses energy at 290 W). Which choice would burn up more
energy, and how much energy (in joules) would it burn? Why does
the more intense exercise burn up less energy than the less intense
exercise?
6.53 .. Magnetar. On December 27, 2004, astronomers observed the greatest flash of light ever recorded from outside the
solar system. It came from the highly magnetic neutron star SGR
1806-20 (a magnetar). During 0.20 s, this star released as much
energy as our sun does in 250,000 years. If P is the average power
output of our sun, what was the average power output (in terms
of P) of this magnetar?
6.54 .. A 20.0-kg rock is sliding on a rough, horizontal surface
at 8.00 m>s and eventually stops due to friction. The coefficient of
kinetic friction between the rock and the surface is 0.200. What
average power is produced by friction as the rock stops?
6.55 . A tandem (two-person) bicycle team must overcome a
force of 165 N to maintain a speed of 9.00 m>s. Find the power
required per rider, assuming that each contributes equally. Express
your answer in watts and in horsepower.
6.56 .. When its 75-kW (100-hp) engine is generating full power,
a small single-engine airplane with mass 700 kg gains altitude
at a rate of 2.5 m>s 1150 m>min, or 500 ft>min2. What fraction of
the engine power is being used to make the airplane climb? (The
remainder is used to overcome the effects of air resistance and of
inefficiencies in the propeller and engine.)
6.57 .. Working Like a Horse. Your job is to lift 30-kg crates
a vertical distance of 0.90 m from the ground onto the bed of a
truck. How many crates would you have to load onto the truck in
1 minute (a) for the average power output you use to lift the crates
to equal 0.50 hp; (b) for an average power output of 100 W?
6.58 .. An elevator has mass 600 kg, not including passengers.
The elevator is designed to ascend, at constant speed, a vertical
distance of 20.0 m (five floors) in 16.0 s, and it is driven by a
motor that can provide up to 40 hp to the elevator. What is the
maximum number of passengers that can ride in the elevator?
Assume that an average passenger has mass 65.0 kg.
6.59 .. A ski tow operates on a 15.0° slope of length 300 m. The
rope moves at 12.0 km>h and provides power for 50 riders at one
time, with an average mass per rider of 70.0 kg. Estimate the
power required to operate the tow.
198
Chapter 6 Work and Kinetic energy
.
S
You are applying a constant horizontal force F =
1- 8.00 N2dn + 13.00 N2en to a crate that is sliding on a factory
floor. At the instant that the velocity of the crate is
S
v = 13.20 m>s2dn + 12.20 m>s2en, what is the instantaneous power
supplied by this force?
6.61 . BIO While hovering, a typical flying insect applies an average force equal to twice its weight during each downward stroke.
Take the mass of the insect to be 10 g, and assume the wings move
an average downward distance of 1.0 cm during each stroke.
Assuming 100 downward strokes per second, estimate the average
power output of the insect.
6.60
probLeMs
... CALC A balky cow is leaving the barn as you try harder
and harder to push her back in. In coordinates with the origin at
the barn door, the cow walks from x = 0 to x = 6.9 m as you
apply a force with x-component Fx = - 320.0 N + 13.0 N>m2x4.
How much work does the force you apply do on the cow during
this displacement?
6.63 . A luggage handler pulls a 20.0-kg suitcase
up a ramp inS
clined at 32.0° above the horizontal by a force F of magnitude
160 N that acts parallel to the ramp. The coefficient of kinetic
friction between the ramp and the incline is mk = 0.300. If the
suitcase travels 3.80 Sm along the ramp, calculate (a) the work done
on the suitcase by F; (b) the work done on the suitcase by the
gravitational force; (c) the work done on the suitcase by the normal force; (d) the work done on the suitcase by the friction force;
(e) the total work done on the suitcase. (f) If the speed of the suitcase is zero at the bottom of the ramp, what is its speed after it has
traveled 3.80 m along the ramp?
6.64 . BIO Chin-ups. While doing a chin-up, a man lifts his
body 0.40 m. (a) How much work must the man do per kilogram
of body mass? (b) The muscles involved in doing a chin-up can
generate about 70 J of work per kilogram of muscle mass. If the
man can just barely do a 0.40-m chin-up, what percentage of his
body’s mass do these muscles constitute? (For comparison, the
total percentage of muscle in a typical 70-kg man with 14% body
fat is about 43%.) (c) Repeat part (b) for the man’s young son, who
has arms half as long as his father’s but whose muscles can also
generate 70 J of work per kilogram of muscle mass. (d) Adults and
children have about the same percentage of muscle in their bodies.
Explain why children can commonly do chin-ups more easily than
their fathers.
6.65 ... Consider the blocks in Exercise 6.7 as they move 75.0 cm.
Find the total work done on each one (a) if there is no friction
between the table and the 20.0-N block, and (b) if ms = 0.500
and mk = 0.325 between the table and the 20.0-N block.
6.66 .. A 5.00-kg package slides 2.80 m down a long ramp that
is inclined at 24.0° below the horizontal. The coefficient of kinetic
friction between the package and the ramp is mk = 0.310.
Calculate (a) the work done on the package by friction; (b) the work
done on the package by gravity; (c) the work done on the package
by the normal force; (d) the total work done on the package. (e) If
the package has a speed of 2.20 m>s at the top of the ramp, what is
its speed after it has slid 2.80 m down the ramp?
6.67 .. CP BIO Whiplash Injuries. When a car is hit from
behind, its passengers undergo sudden forward acceleration,
which can cause a severe neck injury known as whiplash. During
normal acceleration, the neck muscles play a large role in accelerating the head so that the bones are not injured. But during a very
6.62
sudden acceleration, the muscles do not react immediately because
they are flexible; most of the accelerating force is provided by the
neck bones. Experiments have shown that these bones will fracture if they absorb more than 8.0 J of energy. (a) If a car waiting at
a stoplight is rear-ended in a collision that lasts for 10.0 ms, what
is the greatest speed this car and its driver can reach without
breaking neck bones if the driver’s head has a mass of 5.0 kg
(which is about right for a 70-kg person)? Express your answer in
m>s and in mi>h. (b) What is the acceleration of the passengers
during the collision in part (a), and how large a force is acting to
accelerate their heads? Express the acceleration in m>s2 and in g’s.
6.68 .. CALC A net force along the x-axis that has x-component
Fx = - 12.0 N + 10.300 N>m22x 2 is applied to a 5.00-kg object
that is initially at the origin and moving in the - x-direction with a
speed of 6.00 m>s. What is the speed of the object when it reaches
the point x = 5.00 m?
6.69 . CALC Varying Coefficient of Friction. A box is sliding
with a speed of 4.50 m>s on a horizontal surface when, at point P, it
encounters a rough section. The coefficient of friction there is not
constant; it starts at 0.100 at P and increases linearly with distance
past P, reaching a value of 0.600 at 12.5 m past point P. (a) Use the
work–energy theorem to find how far this box slides before stopping. (b) What is the coefficient of friction at the stopping point?
(c) How far would the box have slid if the friction coefficient
didn’t increase but instead had the constant value of 0.100?
6.70 .. CALC Consider a spring that does not obey Hooke’s law
very faithfully. One end of the spring is fixed. To keep the spring
stretched or compressed an amount x, a force along the x-axis with
x-component Fx = kx - bx 2 + cx 3 must be applied to the free
end. Here k = 100 N>m, b = 700 N>m2, and c = 12,000 N>m3.
Note that x 7 0 when the spring is stretched and x 6 0 when it is
compressed. How much work must be done (a) to stretch this
spring by 0.050 m from its unstretched length? (b) To compress
this spring by 0.050 m from its unstretched length? (c) Is it easier
to stretch or compress this spring? Explain why in terms of the
dependence of Fx on x. (Many real springs behave qualitatively in
the same way.)
6.71 .. CP A small block with Figure P6.71
a mass of 0.0600 kg is attached
to a cord passing through a hole
in a frictionless, horizontal surface (Fig. P6.71). The block is
originally revolving at a distance of 0.40 m from the hole
with a speed of 0.70 m>s. The
cord is then pulled from below,
shortening the radius of the circle in which the block revolves
to 0.10 m. At this new distance, the speed of the block is 2.80 m>s.
(a) What is the tension in the cord in the original situation, when
the block has speed v = 0.70 m>s? (b) What is the tension in the
cord in the final situation, when the block has speed v = 2.80 m>s?
(c) How much work was done by the person who pulled on the cord?
6.72 .. CALC Proton Bombardment. A proton with mass
1.67 * 10-27 kg is propelled at an initial speed of 3.00 * 105 m>s
directly toward a uranium nucleus 5.00 m away. The proton is
repelled by the uranium nucleus with a force of magnitude
F = a>x 2 , where x is the separation between the two objects
and a = 2.12 * 10-26 N # m2. Assume that the uranium nucleus
remains at rest. (a) What is the speed of the proton when it is
8.00 * 10-10 m from the uranium nucleus? (b) As the proton
problems
approaches the uranium nucleus, the repulsive force slows down
the proton until it comes momentarily to rest, after which the proton moves away from the uranium nucleus. How close to the uranium nucleus does the proton get? (c) What is the speed of the
proton when it is again 5.00 m away from the uranium nucleus?
6.73 .. You are asked to design spring bumpers for the walls of
a parking garage. A freely rolling 1200-kg car moving at 0.65 m>s
is to compress the spring no more than 0.090 m before stopping.
What should be the force constant of the spring? Assume that the
spring has negligible mass.
6.74 .. You and your bicycle have combined mass 80.0 kg.
When you reach the base of a bridge, you are traveling along the
road at 5.00 m>s (Fig. P6.74). At the top of the bridge, you have
climbed a vertical distance of 5.20 m and slowed to 1.50 m>s.
Ignore work done by friction and any inefficiency in the bike or
your legs. (a) What is the total work done on you and your bicycle
when you go from the base to the top of the bridge? (b) How much
work have you done with the force you apply to the pedals?
Figure P6.74
m = 80.0 kg
5.20 m
6.75
...
A 2.50-kg textbook is forced against a horizontal spring
of negligible mass and force constant 250 N>m, compressing the
spring a distance of 0.250 m. When released, the textbook slides
on a horizontal tabletop with coefficient of kinetic friction
mk = 0.30. Use the work–energy theorem to find how far the textbook moves from its initial position before it comes to rest.
6.76 .. The spring of a spring gun has force constant k =
400 N>m and negligible mass. The spring is compressed 6.00 cm,
and a ball with mass 0.0300 kg is placed in the horizontal barrel
against the compressed spring. The spring is then released, and
the ball is propelled out the barrel of the gun. The barrel is 6.00 cm
long, so the ball leaves the barrel at the same point that it loses
contact with the spring. The gun is held so that the barrel is horizontal. (a) Calculate the speed with which the ball leaves the barrel if you can ignore friction. (b) Calculate the speed of the ball as
it leaves the barrel if a constant resisting force of 6.00 N acts on
the ball as it moves along the barrel. (c) For the situation in part
(b), at what position along the barrel does the ball have the greatest speed, and what is that speed? (In this case, the maximum
speed does not occur at the end of the barrel.)
6.77 .. One end of a horizontal spring with force constant
130.0 N>m is attached to a vertical wall. A 4.00-kg block sitting
on the floor is placed against the spring. The coefficient of kinetic
friction between Sthe block and the
floor is mk = 0.400. You apply
S
a constant force F to the block. F has magnitude F = 82.0 N and
is directed toward the wall. At the instant that the spring is compressed 80.0 cm, what are (a) the speed of the block, and (b) the
magnitude and direction of the block’s acceleration?
199
6.78 .. One end of a horizontal spring with force constant
76.0 N>m is attached to a vertical post. A 2.00-kg block of frictionless ice is attached to the other end and rests on the floor. The
spring is initially neither stretched nor compressed. A constant
horizontal force of 54.0 N is then applied to the block, in the direction away from the post. (a) What is the speed of the block
when the spring is stretched 0.400 m? (b) At that instant, what are
the magnitude and direction of the acceleration of the block?
6.79 . A 5.00-kg block is
moving at v0 = 6.00 m>s along
Figure P6.79
a frictionless, horizontal surv0 = 6.00 m>s
face toward a spring with force
k = 500 N>m
constant k = 500 N>m that is
5.00
attached to a wall (Fig. P6.79).
kg
The spring has negligible mass.
(a) Find the maximum distance
the spring will be compressed. (b) If the spring is to compress
by no more than 0.150 m, what should be the maximum value
of v0?
6.80 ... A physics professor is pushed up a ramp inclined upward at 30.0° above the horizontal as she sits in her desk chair,
which slides on frictionless rollers. The combined mass of the professor and chair is 85.0 kg. She is pushed 2.50 m along the incline
by a group of students who together exert a constant horizontal
force of 600 N. The professor’s speed at the bottom of the ramp is
2.00 m>s. Use the work–energy theorem to find her speed at the
top of the ramp.
6.81 .. Consider the system
shown in Fig. P6.81. The rope Figure P6.81
and pulley have negligible
8.00 kg
mass, and the pulley is frictionless. Initially the 6.00-kg block
is moving downward and the
8.00-kg block is moving to the
right, both with a speed of
0.900 m>s. The blocks come to
6.00 kg
rest after moving 2.00 m. Use
the work–energy theorem to
calculate the coefficient of kinetic friction between the 8.00-kg
block and the tabletop.
6.82 .. Consider the system shown in Fig. P6.81. The rope and
pulley have negligible mass, and the pulley is frictionless. The coefficient of kinetic friction between the 8.00-kg block and the tabletop is mk = 0.250. The blocks are released from rest. Use
energy methods to calculate the speed of the 6.00-kg block after it
has descended 1.50 m.
6.83 .. On an essentially frictionless, horizontal ice rink, a
skater moving at 3.0 m>s encounters a rough patch that reduces
her speed to 1.65 m>s due to a friction force that is 25% of her
weight. Use the work–energy theorem to find the length of this
rough patch.
6.84 .. bio All birds, independent of their size, must maintain a
power output of 10–25 watts per kilogram of body mass in order
to fly by flapping their wings. (a) The Andean giant hummingbird
(Patagona gigas) has mass 70 g and flaps its wings 10 times per
second while hovering. Estimate the amount of work done by such
a hummingbird in each wingbeat. (b) A 70-kg athlete can maintain a power output of 1.4 kW for no more than a few seconds; the
steady power output of a typical athlete is only 500 W or so. Is it
possible for a human-powered aircraft to fly for extended periods
by flapping its wings? Explain.
200
Chapter 6 Work and Kinetic energy
6.85 .. A pump is required to lift 800 kg of water (about 210 gallons) per minute from a well 14.0 m deep and eject it with a
speed of 18.0 m>s. (a) How much work is done per minute in lifting the water? (b) How much work is done in giving the water the
kinetic energy it has when ejected? (c) What must be the power
output of the pump?
6.86 ... The Grand Coulee Dam is 1270 m long and 170 m high.
The electrical power output from generators at its base is approximately 2000 MW. How many cubic meters of water must flow
from the top of the dam per second to produce this amount of
power if 92% of the work done on the water by gravity is converted to electrical energy? (Each cubic meter of water has a mass
of 1000 kg.)
6.87 ... A physics student spends part of her day walking between classes or for recreation, during which time she expends
energy at an average rate of 280 W. The remainder of the day she
is sitting in class, studying, or resting; during these activities, she
expends energy at an average rate of 100 W. If she expends a total
of 1.1 * 107 J of energy in a 24-hour day, how much of the day
did she spend walking?
6.88 . CALC An
object has several forces acting on it. One of
S
these forces is F = axydn, a force in the x-direction whose magnitude depends on the position of the object, with a = 2.50 N>m2.
Calculate the work done on the object by this force for the following displacements of the object: (a) The object starts at the
point 1x = 0, y = 3.00 m2 and moves parallel to the x-axis to the
point 1x = 2.00 m, y = 3.00 m2. (b) The object starts at the point
1x = 2.00 m, y = 02 and moves in the y-direction to the point
1x = 2.00 m, y = 3.00 m2. (c) The object starts at the origin and
moves on the line y = 1.5x to the point 1x = 2.00 m, y = 3.00 m2.
6.89 . bio Power of the Human Heart. The human heart is
a powerful and extremely reliable pump. Each day it takes in and
discharges about 7500 L of blood. Assume that the work done by
the heart is equal to the work required to lift this amount of blood
a height equal to that of the average American woman (1.63 m).
The density (mass per unit volume) of blood is 1.05 * 103 kg>m3.
(a) How much work does the heart do in a day? (b) What is the
heart’s power output in watts?
6.90 .. DATA Figure P6.90 shows the results of measuring the
force F exerted on both ends of a rubber band to stretch it a distance
x from its unstretched position. (Source: www.sciencebuddies.org)
The data points are well fit by the equation F = 33.55x 0.4871,
where F is in newtons and x is in meters. (a) Does this rubber band
obey Hooke’s law over the range of x shown in the graph? Explain.
(b) The stiffness of a spring that obeys Hooke’s law is measured
by the value of its force constant k, where k = F>x. This can be
Figure P6.90
F (N)
written as k = dF>dx to emphasize the quantities that are changing. Define keff = dF>dx and calculate keff as a function of x for
this rubber band. For a spring that obeys Hooke’s law, keff is constant, independent of x. Does the stiffness of this band, as measured by keff , increase or decrease as x is increased, within the
range of the data? (c) How much work must be done to stretch the
rubber band from x = 0 to x = 0.0400 m? From x = 0.0400 m to
x = 0.0800 m? (d) One end of the rubber band is attached to a
stationary vertical rod, and the band is stretched horizontally
0.0800 m from its unstretched length. A 0.300-kg object on a horizontal, frictionless surface is attached to the free end of the rubber band and released from rest. What is the speed of the object
after it has traveled 0.0400 m?
6.91 ... DATA In a physics lab experiment,
one end of a horizontal spring that obeys d 1m2
v 1m , s 2
Hooke’s law is attached to a wall. The spring
0
0
is compressed 0.400 m, and a block with
0.05
0.85
mass 0.300 kg is attached to it. The spring is
0.10
1.11
then released, and the block moves along a
horizontal surface. Electronic sensors mea0.15
1.24
sure the speed v of the block after it has trav0.25
1.26
eled a distance d from its initial position
0.30
1.14
against the compressed spring. The measured
0.35
0.90
values are listed in the table. (a) The data
0.40
0.36
show that the speed v of the block increases
and then decreases as the spring returns to its unstretched length.
Explain why this happens, in terms of the work done on the block
by the forces that act on it. (b) Use the work–energy theorem to
derive an expression for v 2 in terms of d. (c) Use a computer graphing program (for example, Excel or Matlab) to graph the data as v 2
(vertical axis) versus d (horizontal axis). The equation that you derived in part (b) should show that v 2 is a quadratic function of d, so,
in your graph, fit the data by a second-order polynomial (quadratic)
and have the graphing program display the equation for this trendline. Use that equation to find the block’s maximum speed v and
the value of d at which this speed occurs. (d) By comparing the
equation from the graphing program to the formula you derived in
part (b), calculate the force constant k for the spring and the coefficient of kinetic friction for the friction force that the surface
exerts on the block.
6.92 .. DATA For a physics lab experiment, four classmates run
up the stairs from the basement to the top floor of their physics
building—a vertical distance of 16.0 m. The classmates and their
masses are: Tatiana, 50.2 kg; Bill, 68.2 kg; Ricardo, 81.8 kg;
and Melanie, 59.1 kg. The time it takes each of them is shown in
Fig. P6.92. (a) Considering only the work done against gravity,
which person had the largest average power output? The smallest?
(b) Chang is very fit and has mass 62.3 kg. If his average power
output is 1.00 hp, how many seconds does it take him to run up the
stairs?
12
Figure P6.92
10
Time (s)
60
8
50
6
40
4
30
20
2
0
0.02
0.04
0.06
0.08
0.1
x (m)
10
0
Tatiana
Bill
Ricardo
Melanie
Passage Problems
challEngE ProblEMs
6.93 ... CALC A Spring with Mass. We usually ignore the
kinetic energy of the moving coils of a spring, but let’s try to get a
reasonable approximation to this. Consider a spring of mass M,
equilibrium length L 0, and force constant k. The work done to
stretch or compress the spring by a distance L is 12 kX 2, where
X = L - L 0. Consider a spring, as described above, that has one
end fixed and the other end moving with speed v. Assume that
the speed of points along the length of the spring varies linearly
with distance l from the fixed end. Assume also that the mass M
of the spring is distributed uniformly along the length of the spring.
(a) Calculate the kinetic energy of the spring in terms of M and v.
(Hint: Divide the spring into pieces of length dl; find the speed of
each piece in terms of l, v, and L; find the mass of each piece in
terms of dl, M, and L; and integrate from 0 to L. The result is not
1
2
2 Mv , since not all of the spring moves with the same speed.) In
a spring gun, a spring of mass 0.243 kg and force constant
3200 N>m is compressed 2.50 cm from its unstretched length.
When the trigger is pulled, the spring pushes horizontally on a
0.053-kg ball. The work done by friction is negligible. Calculate
the ball’s speed when the spring reaches its uncompressed length
(b) ignoring the mass of the spring and (c) including, using the
results of part (a), the mass of the spring. (d) In part (c), what is the
final kinetic energy of the ball and of the spring?
6.94 ... CALC An airplane in flight is subject to an air resistance
force proportional to the square of its speed v. But there is an additional resistive force because the airplane has wings. Air flowing
over the wings is pushed down and slightly forward, so from
Newton’s third law the air exerts a force on the wings and airplane
that is up and slightly backward (Fig. P6.94). The upward force is
the lift force that keeps the airplane aloft, and the backward force
is called induced drag. At flying speeds, induced drag is inversely
proportional to v 2, so the total air resistance force can be expressed
by Fair = av 2 + b>v 2, where a and b are positive constants that
depend on the shape and size of the airplane and the density of the
air. For a Cessna 150, a small single-engine airplane, a =
0.30 N # s2>m2 and b = 3.5 * 105 N # m2>s2. In steady flight, the
engine must provide a forward force that exactly balances the air
resistance force. (a) Calculate the speed 1in km>h2 at which this
airplane will have the maximum range (that is, travel the greatest
distance) for a given quantity of fuel. (b) Calculate the speed (in
km>h) for which the airplane will have the maximum endurance
(that is, remain in the air the longest time).
201
Figure P6.94
Induced drag
Lift
Force of air
on wings
PassagE ProblEMs
BIO EnErgy of locomotion. On flat ground, a 70-kg person requires about 300 W of metabolic power to walk at a steady
pace of 5.0 km>h11.4 m>s2. Using the same metabolic power output, that person can bicycle over the same ground at 15 km>h.
6.95 Based on the given data, how does the energy used in biking 1 km compare with that used in walking 1 km? Biking takes
(a) 13 of the energy of walking the same distance; (b) the same energy as walking the same distance; (c) 3 times the energy of walking the same distance; (d) 9 times the energy of walking the same
distance.
6.96 A 70-kg person walks at a steady pace of 5.0 km>h on a
treadmill at a 5.0% grade. (That is, the vertical distance covered is
5.0% of the horizontal distance covered.) If we assume the metabolic power required is equal to that required for walking on a
flat surface plus the rate of doing work for the vertical climb,
how much power is required? (a) 300 W; (b) 315 W; (c) 350 W;
(d) 370 W.
6.97 How many times greater is the kinetic energy of the person
when biking than when walking? Ignore the mass of the bike.
(a) 1.7; (b) 3; (c) 6; (d) 9.
202
Chapter 6 Work and Kinetic energy
Answers
Chapter opening Question
?
(ii) The expression for kinetic energy is K = 12 mv 2. If we calculate K for the three balls, we find (i) K = 12 10.145 kg2 *
120.0 m>s22 = 29.0 kg # m2>s2 = 29.0 J, (ii) K = 12 10.0145 kg2 *
1200 m>s22 = 290 J, and (iii) K = 12 11.45 kg212.00 m>s22 =
2.90 J. The smaller ball has the least mass of all three, but it also
has the greatest speed and so the most kinetic energy. Since kinetic
energy is a scalar, it does not depend on the direction of motion.
test your understanding Questions
6.1 (iii) The electron has constant velocity, so its acceleration
is zero and (by Newton’s second law) the net force on the electron is also zero. Therefore the total work done by all the forces
(equal to the work done by the net force) must be zero as well.
The individual forces may do nonzero work, but that’s not what
the question asks.
6.2 (iv), (i), (iii), (ii) Body (i) has kinetic energy K = 12 mv 2 =
1
2
2 12.0 kg215.0 m>s2 = 25 J. Body (ii) had zero kinetic energy
initially and then had 30 J of work done on it, so its final kinetic
energy is K2 = K1 + W = 0 + 30 J = 30 J. Body (iii) had initial
kinetic energy K1 = 12 mv 12 = 12 11.0 kg214.0 m>s22 = 8.0 J and
then had 20 J of work done on it, so its final kinetic energy is
K2 = K1 + W = 8.0 J + 20 J = 28 J. Body (iv) had initial kinetic
energy K1 = 12 mv 12 = 12 12.0 kg2110 m>s22 = 100 J; when it did
80 J of work on another body, the other body did -80 J of work
on body (iv), so the final kinetic energy of body (iv) is K2 =
K1 + W = 100 J + 1- 80 J2 = 20 J.
6.3 (a) (iii), (b) (iii) At any point during the pendulum bob’s
motion, both the tension force and the weight act perpendicular to
the motion—that
is, perpendicular to an infinitesimal displaceS
S
ment dl of the bob. (In Fig. 5.32b, the displacement dl would
be directed outward from the plane of the free-body diagram.)
Hence for either
force the scalar product inside the integral in
S
S
Eq. (6.14) is F ~ dl = 0, and the work done along anyS partS of the
circular path (including a complete circle) is W = 1 F ~ dl = 0.
6.4 (v) The airliner has a constant horizontal velocity, so the
net horizontal force on it must be zero. Hence the backward drag
force must have the same magnitude as the forward force due
to the combined thrust of the four engines. This means that the
drag force must do negative work on the airplane at the same rate
that the combined thrust force does positive work. The combined
thrust does work at a rate of 41108,000 hp2 = 432,000 hp, so the
drag force must do work at a rate of -432,000 hp.
Bridging Problem
(a) v1 =
2ax 13
2
1mgx1 - 13 ax 132 = 2gx1 Am
3m
B
(b) P = - Fspring - 1v1 = - ax 12
(c) x2 =
3mg
A a
(d) No
B
2gx1 -
2ax 13
3m
?
As this sandhill crane
(Grus canadensis) glides
in to a landing, it descends
along a straight-line path at
a constant speed. During the
glide, what happens to the
mechanical energy (the sum of
kinetic energy and gravitational
potential energy)? (i) It stays
the same; (ii) it increases due
to the effect of gravity; (iii) it
increases due to the effect of
the air; (iv) it decreases due
to the effect of gravity; (v) it
decreases due to the effect of
the air.
7
Potential energy
and energy
Conservation
learninG Goals
Looking forward at …
7.1 How to use the concept of gravitational
7.2
7.3
7.4
7.5
potential energy in problems that involve
vertical motion.
How to use the concept of elastic potential
energy in problems that involve a moving
body attached to a stretched or compressed
spring.
The distinction between conservative and
nonconservative forces, and how to solve
problems in which both kinds of forces act
on a moving body.
How to calculate the properties of a conservative force if you know the corresponding
potential-energy function.
How to use energy diagrams to understand
how an object moves in a straight line
under the influence of a conservative force.
Looking back at …
5.3 Kinetic friction and fluid resistance.
5.4 Dynamics of circular motion.
6.1, 6.2 Work and the work–energy theorem.
6.3 Work done by an ideal spring.
W
hen a diver jumps off a high board into a swimming pool, she hits
the water moving pretty fast, with a lot of kinetic energy—energy of
motion. Where does that energy come from? The answer we learned in
Chapter 6 was that the gravitational force does work on the diver as she falls, and
her kinetic energy increases by an amount equal to the work done.
However, there’s a useful alternative way to think about work and kinetic
energy. This new approach uses the idea of potential energy, which is associated
with the position of a system rather than with its motion. In this approach, there
is gravitational potential energy even when the diver is at rest on the high board.
As she falls, this potential energy is transformed into her kinetic energy.
If the diver bounces on the end of the board before she jumps, the bent board
stores a second kind of potential energy called elastic potential energy. We’ll discuss elastic potential energy of simple systems such as a stretched or compressed
spring. (An important third kind of potential energy is associated with the forces
between electrically charged objects. We’ll return to this in Chapter 23.)
We will prove that in some cases the sum of a system’s kinetic and potential
energies, called the total mechanical energy of the system, is constant during
the motion of the system. This will lead us to the general statement of the law of
conservation of energy, one of the most fundamental principles in all of science.
7.1 Gravitational Potential enerGy
In many situations it seems as though energy has been stored in a system, to be
recovered later. For example, you must do work to lift a heavy stone over your
head. It seems reasonable that in hoisting the stone into the air you are storing
energy in the system, energy that is later converted into kinetic energy when you
let the stone fall.
203
204
Chapter 7 potential energy and energy Conservation
7.1 The greater the height of a basketball,
the greater the associated gravitational
potential energy. As the basketball
descends, gravitational potential energy
is converted to kinetic energy and the
basketball’s speed increases.
7.2 When a body moves vertically from
an initial height y1 to a final height y2 , the
S
gravitational force w does work and the
gravitational potential energy changes.
(a) A body moves downward
S
S
Fother
S
s
y1
Displacement s
is downward and
y decreases
(y1 7 y2),
S
so w does
y2 - y1
positive work
and gravitational
potential energy
decreases:
∆Ugrav 6 0.
y2
S
S
w = mg
O
(b) A body moves upward
Displacement s
is upward and
y increases
(y1 6 y2),
S
so w does
S
S
w = mg
y2 - y1 negative work
and gravitational
y2
potential energy
increases:
∆Ugrav 7 0.
y1
O
(7.1)
This expression also gives the correct work when the body moves upward and y2
is greater than y1 (Fig. 7.2b). In that case the quantity 1y1 - y22 is negative, and
Wgrav is negative because the weight and displacement are opposite in direction.
Equation (7.1) shows that we can express Wgrav in terms of the values of the
quantity mgy at the beginning and end of the displacement. This quantity is
called the gravitational potential energy, Ugrav :
Gravitational potential energy
associated with a particle
Ugrav = mgy
Mass of particle
Vertical coordinate of particle
(y increases if particle
moves upward)
(7.2)
Acceleration due to gravity
Its initial value is Ugrav, 1 = mgy1 and its final value is Ugrav, 2 = mgy2 . The change
in Ugrav is the final value minus the initial value, or ∆Ugrav = Ugrav, 2 - Ugrav, 1 .
Using Eq. (7.2), we can rewrite Eq. (7.1) for the work done by the gravitational
force during the displacement from y1 to y2:
Wgrav = Ugrav, 1 - Ugrav, 2 = -1Ugrav, 2 - Ugrav, 12 = - ∆Ugrav
Work done by the gravitational
force on a particle ...
S
s
Wgrav = Fs = w1y1 - y22 = mgy1 - mgy2
or
S
Fother
S
This example points to the idea of an energy associated with the position of
bodies in a system. This kind of energy is a measure of the potential or possibility
for work to be done; if you raise a stone into the air, there is a potential for the
gravitational force to do work on it, but only if you allow the stone to fall to the
ground. For this reason, energy associated with position is called potential energy.
The potential energy associated with a body’s weight and its height above the
ground is called gravitational potential energy (Fig. 7.1).
We now have two ways to describe what happens when a body falls without air
resistance. One way, which we learned in Chapter 6, is to say that a falling body’s
kinetic energy increases because the force of the earth’s gravity does work on the
body. The other way is to say that the kinetic energy increases as the gravitational potential energy decreases. Later in this section we’ll use the work–energy
theorem to show that these two descriptions are equivalent.
Let’s derive the expression for gravitational potential energy. Suppose a body
with mass m moves along the (vertical) y-axis, as in Fig. 7.2. The forces acting
on it are its weight, with magnitude w = mg, and possibly
some other forces; we
S
call the vector sum (resultant) of all the other forces Fother . We’ll assume that the
body stays close enough to the earth’s surface that the weight is constant. (We’ll
find in Chapter 13 that weight decreases with altitude.) We want to find the work
done by the weight when the body moves downward from a height y1 above the
origin to a lower height y2 (Fig. 7.2a). The weight and displacement are in the
same direction, so the work Wgrav done on the body by its weight is positive:
... equals the negative of the change in
the gravitational potential energy.
Wgrav = mgy1 - mgy2 = Ugrav, 1 - Ugrav, 2 = -∆Ugrav
Mass of
particle
Acceleration
due to gravity
(7.3)
Initial and final vertical
coordinates of particle
The negative sign in front of ∆Ugrav is essential. When the body moves up,
y increases, the work done by the gravitational force is negative, and the gravitational potential energy increases 1∆Ugrav 7 02. When the body moves down,
y decreases, the gravitational force does positive work, and the gravitational
potential energy decreases 1∆Ugrav 6 02. It’s like drawing money out of the
bank (decreasing Ugrav) and spending it (doing positive work). The unit of potential
energy is the joule (J), the same unit as is used for work.
7.1 Gravitational Potential Energy
Caution To what body does gravitational potential energy “belong”? It is not correct
to call Ugrav = mgy the “gravitational potential energy of the body.” The reason is that
Ugrav is a shared property of the body and the earth. The value of Ugrav increases if the
earth stays fixed and the body moves upward, away from the earth; it also increases
if the body stays fixed and the earth is moved away from it. Notice that the formula
Ugrav = mgy involves characteristics of both the body (its mass m) and the earth (the
value of g). ❙
Conservation of Mechanical Energy
(Gravitational Forces only)
205
BIO application Converting
Gravitational Potential Energy
to Kinetic Energy When a kingfisher
(Alcedo atthis) spots a tasty fish, the bird dives
from its perch with its wings tucked in to minimize air resistance. Effectively the only force
acting on the diving kingfisher is the force of
gravity, so mechanical energy is conserved:
The gravitational potential energy lost as the
kingfisher descends is converted into the
bird’s kinetic energy.
To see what gravitational potential
energy is good for, suppose a body’s weight
S
is the only force acting on it, so Fother = 0. The body is then falling freely with
no air resistance and can be moving either up or down. Let its speed at point y1
be v1 and let its speed at y2 be v2 . The work–energy theorem, Eq. (6.6), says that
the total work done on the body equals the change in the body’s kinetic energy:
Wtot = ∆K = K2 - K1 . If gravity is the only force that acts, then from Eq. (7.3),
Wtot = Wgrav = - ∆Ugrav = Ugrav, 1 - Ugrav, 2 . Putting these together, we get
∆K = - ∆Ugrav
or
K2 - K1 = Ugrav, 1 - Ugrav, 2
which we can rewrite as
If only the gravitational force does work, total mechanical energy is conserved:
Initial gravitational potential energy
Initial kinetic energy
Ugrav, 1 = mgy1
K1 = 12 mv12
K1 + Ugrav, 1 = K2 + Ugrav, 2
Final kinetic energy
K2 = 12 mv22
(7.4)
Final gravitational potential energy
Ugrav, 2 = mgy2
The sum K + Ugrav of kinetic and potential energies is called E, the total
mechanical energy of the system. By “system” we mean the body of mass m
and the earth considered together, because gravitational potential energy U
is a shared property of both bodies. Then E 1 = K1 + Ugrav, 1 is the total mechani­
cal energy at y1 and E 2 = K2 + Ugrav, 2 is the total mechanical energy at y2 .
Equation (7.4) says that when the body’s weight is the only force doing work on it,
E 1 = E 2 . That is, E is constant; it has the same value at y1 and y2 . But since
positions y1 and y2 are arbitrary points in the motion of the body, the total
mechanical energy E has the same value at all points during the motion:
E = K + Ugrav = constant
Demo
(if only gravity does work)
A quantity that always has the same value is called a conserved quantity. When
only the force of gravity does work, the total mechanical energy is constant—
that is, it is conserved (Fig. 7.3). This is our first example of the conservation of
mechanical energy.
Moving up:
• K decreases.
• Ugrav increases.
• E = K + Ugrav
stays the same.
Moving down:
• K increases.
• Ugrav decreases.
• E = K + Ugrav
stays the same.
S
S
w = mg
7.3 While this athlete is in midair, only
gravity does work on him (if we neglect
the minor effects of air resistance).
Mechanical energy E—the sum of kinetic
and gravitational potential energy—is
conserved.
206
Chapter 7 potential energy and energy Conservation
When we throw a ball into the air, its speed decreases on the way up as
kinetic energy is converted to potential energy: ∆K 6 0 and ∆Ugrav 7 0. On
the way back down, potential energy is converted back to kinetic energy and
the ball’s speed increases: ∆K 7 0 and ∆Ugrav 6 0. But the total mechanical energy (kinetic plus potential) is the same at every point in the motion,
provided that no force other than gravity does work on the ball (that is, air
resistance must be negligible). It’s still true that the gravitational force does
work on the body as it moves up or down, but we no longer have to calculate
work directly; keeping track of changes in the value of Ugrav takes care of this
completely.
Equation (7.4) is also valid if forces other than gravity are present but do not
do work. We’ll see a situation of this kind later, in Example 7.4.
Caution Choose “zero height” to be wherever you like When working with gravitational potential energy, we may choose any height to be y = 0. If we shift the origin for y,
the values of y1 and y2 change, as do the values of Ugrav, 1 and Ugrav, 2 . But this shift has no
effect on the difference in height y2 - y1 or on the difference in gravitational potential
energy Ugrav, 2 - Ugrav, 1 = mg1y2 - y12. As Example 7.1 shows, the physically significant
quantity is not the value of Ugrav at a particular point but the difference in Ugrav between
two points. We can define Ugrav to be zero at whatever point we choose. ❙
HEigHt of a basEball from EnErgy ConsErvation
You throw a 0.145-kg baseball straight up, giving it an initial
velocity of magnitude 20.0 m>s. Find how high it goes, ignoring
air resistance.
Solution
identify and Set up: After the ball leaves your hand, only grav-
ity does work on it. Hence mechanical energy is conserved, and
we can use Eq. (7.4). We take point 1 to be where the ball leaves
your hand and point 2 to be where it reaches its maximum height.
As in Fig. 7.2, we take the positive y-direction to be upward. The
ball’s speed at point 1 is v1 = 20.0 m>s; at its maximum height it
is instantaneously at rest, so v2 = 0. We take the origin at point 1,
so y1 = 0 (Fig. 7.4). Our target variable, the distance the ball
moves vertically between the two points, is the displacement
y2 - y1 = y2 - 0 = y2.
7.4 After a baseball leaves your hand, mechanical energy
E = K + U is conserved.
Energy at y2
After the ball leaves your
hand, the only force
acting on it is gravity ...
zero
v2 = 0
y2
E = K + Ugrav
... so the mechanical energy
E = K + U stays constant.
Energy at y1
v1 = 20.0 m>s
y1 = 0
zero
m = 0.145 kg
Solution
ExamplE 7.1
E = K + Ugrav
exeCute: We have y1 = 0, Ugrav, 1 = mgy1 = 0, and K2 =
1
2
2 mv 2
= 0. Then Eq. (7.4), K1 + Ugrav, 1 = K2 + Ugrav, 2, becomes
K1 = Ugrav, 2
As the energy bar graphs in Fig. 7.4 show, this equation says that
the kinetic energy of the ball at point 1 is completely converted to
gravitational potential energy at point 2. We substitute K1 = 12 mv 12
and Ugrav, 2 = mgy2 and solve for y2:
1
2
2 mv 1
= mgy2
y2 =
120.0 m>s22
v 12
=
= 20.4 m
2g
219.80 m>s22
evaluate: As a check, use the given value of v1 and our result for
y2 to calculate the kinetic energy at point 1 and the gravitational
potential energy at point 2. You should find that these are equal:
K1 = 12 mv 12 = 29.0 J and Ugrav, 2 = mgy2 = 29.0 J. Note that we
could have found the result y2 = v 12>2g by using Eq. (2.13) in the
form v 2y 2 = v 1y 2 - 2g1y2 - y12.
What if we put the origin somewhere else—for example, 5.0 m
below point 1, so that y1 = 5.0 m? Then the total mechanical
energy at point 1 is part kinetic and part potential; at point 2 it’s
still purely potential because v2 = 0. You’ll find that this choice
of origin yields y2 = 25.4 m, but again y2 - y1 = 20.4 m. In
problems like this, you are free to choose the height at which
Ugrav = 0. The physics doesn’t depend on your choice.
7.1 Gravitational Potential Energy
When Forces other than Gravity Do Work
207
S
If other forces act on the body in addition to its weight, then Fother in Fig. 7.2 is
not zero. For the pile driver described in Example 6.4 (Section 6.2), the force
applied by the hoisting cable and the friction
with the vertical guide rails are
S
examples of forces that might be included in Fother . The gravitational work Wgrav
is still given bySEq. (7.3), but the total work Wtot is then the sum of Wgrav and the
work done by Fother . We will call this additional work Wother , so the total work
done by all forces is Wtot = Wgrav + Wother . Equating this to the change in kinetic
energy, we have
Wother + Wgrav = K2 - K1
(7.5)
7.5 As this parachutist moves downward,
the upward force of air resistance does
negative work Wother on him. Hence the
total mechanical energy E = K + U
decreases.
Also, from Eq. (7.3), Wgrav = Ugrav, 1 - Ugrav, 2 , so Eq. (7.5) becomes
Wother + Ugrav, 1 - Ugrav, 2 = K2 - K1
which we can rearrange in the form
K1 + Ugrav, 1 + Wother = K2 + Ugrav, 2
(if forces other than
gravity do work)
(7.6)
We can use the expressions for the various energy terms to rewrite Eq. (7.6):
1
2
2 mv 1
+ mgy1 + Wother = 12 mv 22 + mgy2
(if forces other than
gravity do work)
(7.7)
The meaning of Eqs. (7.6) and (7.7) is this: The work done by all forces other
than the gravitational force equals the change in the total mechanical energy E =
K + Ugrav of the system, where Ugrav is the gravitational potential energy. When
Wother is positive, E increases and K2 + Ugrav, 2 is greater than K1 + Ugrav, 1 .
When Wother is negative, E decreases (Fig. 7.5). In the special case in which no
forces other than the body’s weight do work, Wother = 0. The total mechanical
energy is then constant, and we are back to Eq. (7.4).
Problem-Solving STraTegy 7.1
S
Fother
(air resistance)
S
s
(displacement)
# Fother and s are opposite, so Wother 6 0.
# Hence E = K + Ugrav must decrease.
# The parachutist’s speed remains constant, so
S
S
K is constant.
# The parachutist descends, so Ugrav decreases.
ProblemS USing mechanical energy i
iDEntiFy the relevant concepts: Decide whether the problem
should be solved by energy methods, by using gF = ma directly,
or by a combination of these. The energy approach is best when
the problem involves varying forces or motion along a curved path
(discussed later in this section). If the problem involves elapsed
time, the energy approach is usually not the best choice because it
doesn’t involve time directly.
any forces other than gravity. In Section 7.2 we’ll see that the
work done by an ideal spring can also be expressed as a
change in potential energy.) Sketch a free­body diagram for
each body.
4. List the unknown and known quantities, including the coordi­
nates and velocities at each point. Identify the target variables.
SEt uP the problem using the following steps:
1. When using the energy approach, first identify the initial and
final states (the positions and velocities) of the bodies in ques­
tion. Use the subscript 1 for the initial state and the subscript 2
for the final state. Draw sketches showing these states.
2. Define a coordinate system, and choose the level at which
y = 0. Choose the positive y­direction to be upward. (The
equations in this section require this.)
3. Identify any forces that do work on each body and that cannot
be described in terms of potential energy. (So far, this means
ExECutE the solution: Write expressions for the initial and final
kinetic and potential energies K1, K2, Ugrav, 1, and Ugrav, 2. If no
other forces do work, use Eq. (7.4). If there are other forces that do
work, use Eq. (7.6). Draw bar graphs showing the initial and final
values of K, Ugrav, 1, and E = K + Ugrav. Then solve to find your
target variables.
S
S
EvaluatE your answer: Check whether your answer makes phys­
ical sense. Remember that the gravitational work is included in
∆Ugrav, so do not include it in Wother.
208
Chapter 7 potential energy and energy Conservation
Solution
ExamplE 7.2 Work and EnErgy in THroWing a basEball
In Example 7.1 suppose your hand moves upward by 0.50 m while
you are throwing the ball. The ball leaves your hand with an
upward velocity of 20.0 m>s. (a) Find the magnitude of the force
(assumed constant) that your hand exerts on the ball. (b) Find the
speed of the ball at a point 15.0 m above the point where it leaves
your hand. Ignore air resistance.
solution
the work Wother done by this force. We have
K1 = 0
Ugrav, 1 = mgy1 = 10.145 kg219.80 m>s221- 0.50 m2 = - 0.71 J
K2 = 12 mv 22 = 12 10.145 kg2120.0 m>s22 = 29.0 J
Ugrav, 2 = mgy2 = 10.145 kg219.80 m>s22102 = 0
(Don’t worry that Ugrav, 1 is less than zero; all that matters is the
difference in potential energy from one point to another.) From
Eq. (7.6),
K1 + Ugrav, 1 + Wother = K2 + Ugrav, 2
Wother = 1K2 - K12 + 1Ugrav, 2 - Ugrav, 12
S
= 129.0 J - 02 + 30 - 1- 0.71 J24 = 29.7 J
S
But since F is constant and upward, the work done by F equals the
force magnitude times the displacement: Wother = F1y2 - y12. So
Wother
29.7 J
F =
=
= 59 N
y2 - y1
0.50 m
This is more than 40 times the weight of the ball (1.42 N).
(b) To find v3y, note that between points 2 and 3 only gravity acts on the ball. So between these points mechanical energy is
v3
y3 = 15.0 m
E = K + Ugrav
... so the total mechanical
energy E = K + U
stays constant.
After the ball leaves your
hand, the only force
acting on it is gravity ...
(b)
y
v2 = 20.0 m>s
y2 = 0
zero
exeCute: (a) To determine F, we’ll first use Eq. (7.6) to calculate
(a)
E = K + Ugrav
F
... so the total
As you throw the ball,
mechanical energy
0.50
m
you do positive work
E increases.
Wother on it ...
zero
identiFy and set uP: In Example 7.1 only gravity did work.
Here we must include the nongravitational, “other” work done by
your hand. Figure 7.6 shows a diagram of the situation, including a free-body diagram for the ball while it is being thrown. We
let point 1 be where your hand begins to move, point 2 be where
the ball leaves your hand, and point 3 be where
the ball is 15.0 m
S
above point 2. The nongravitational force F of your hand acts only
between points 1 and 2. Using the same coordinate system as in
Example 7.1, we have y1 = - 0.50 m, y2 = 0, and y3 = 15.0 m.
The ball starts at rest at point 1, so v1 = 0, and the ball’s speed as
it leaves your hand is v2 = 20.0 m>s. Our target variables are (a) the
magnitude F of the force of your hand and (b) the magnitude of
the ball’s velocity v3y at point 3.
7.6 (a) Applying energy ideas to a ball thrown vertically upward.
(b) Free-body diagram for the ball as you throw it.
v1 = 0
y1 = -0.50 m
E = K + Ugrav
w
x
conserved and Wother = 0. From Eq. (7.4), we can solve for K3 and
from that solve for v3y:
K2 + Ugrav, 2 = K3 + Ugrav, 3
Ugrav, 3 = mgy3 = 10.145 kg219.80 m>s22115.0 m2 = 21.3 J
K3 = 1K2 + Ugrav, 22 - Ugrav, 3
Since K3 =
= 129.0 J + 0 J2 - 21.3 J = 7.7 J
1
2
2 mv 3y ,
v3y = {
we find
2K3
B m
= {
217.7 J2
B 0.145 kg
= {10 m>s
The plus-or-minus sign reminds us that the ball passes point 3 on
the way up and again on the way down. The ball’s kinetic energy
K3 = 7.7 J at point 3, and hence its speed at that point, doesn’t depend
on the direction the ball is moving. The velocity v3y is positive
1+ 10 m>s2 when the ball is moving up and negative 1- 10 m>s2
when it is moving down; the speed v3 is 10 m>s in either case.
evaluate: In Example 7.1 we found that the ball reaches a maximum height y = 20.4 m. At that point all of the kinetic energy it
had when it left your hand at y = 0 has been converted to gravitational potential energy. At y = 15.0 m, the ball is about threefourths of the way to its maximum height, so about three-fourths
of its mechanical energy should be in the form of potential energy.
Can you verify this from our results for K3 and Ugrav, 3?
Gravitational Potential energy for Motion
along a Curved Path
In our first two examples the body moved along a straight vertical line. What
happens when the path is slanted or curved (Fig. 7.7a)? The body is acted on by
S
S
the gravitational force w = mg and possibly by other forces whose resultant we
209
7.1 Gravitational potential energy
S
call Fother . To find the work Wgrav done by the gravitational force during this disS
placement, we divide the path into small segments ∆s ; Fig. 7.7b shows a typical
segment. The work done by the gravitational force over this segment is the scalar
product of the force and the displacement. In terms of unit vectors, the force is
S
S
S
w = mg = -mgen and the displacement is ∆s = ∆xdn + ∆yen, so
7.7 Calculating the change in gravitational
potential energy for a displacement along a
curved path.
(a)
S
S
Wgrav = w # ∆s = -mgen # 1∆xdn + ∆yen2 = -mg∆y
S
Fother
y1
The work done by gravity is the same as though the body had been displaced
vertically a distance ∆y, with no horizontal displacement. This is true for every
segment, so the total work done by the gravitational force is -mg multiplied by
the total vertical displacement 1y2 - y12:
S
y2
O
Wgrav = -mg1y2 - y12 = mgy1 - mgy2 = Ugrav, 1 - Ugrav, 2
(b)
This is the same as Eq. (7.1) or (7.3), in which we assumed a purely vertical path.
So even if the path a body follows between two points is curved, the total work
done by the gravitational force depends on only the difference in height between
the two points of the path. This work is unaffected by any horizontal motion that
may occur. So we can use the same expression for gravitational potential energy
whether the body’s path is curved or straight.
The work done by the gravitational
force depends only on the vertical
component of displacement ∆y.
∆x
∆y
S
S
w = mg
S
∆s
In this case
∆y is negative.
Solution
ConCEptual ExamplE 7.3
S
w = mg
EnErgy in projECtilE motion
A batter hits two identical baseballs with the same initial speed
and from the same initial height but at different initial angles.
Prove that both balls have the same speed at any height h if air
resistance can be ignored.
7.8 For the same initial speed and initial height, the speed of a
projectile at a given elevation h is always the same, if we ignore
air resistance.
y
Solution
h
zero
The only force acting on each ball after it is hit is its weight. Hence
the total mechanical energy for each ball is constant. Figure 7.8
shows the trajectories of two balls batted at the same height with
the same initial speed, and thus the same total mechanical energy,
but with different initial angles. At all points at the same height
the potential energy is the same. Thus the kinetic energy at this
height must be the same for both balls, and the speeds are the
same.
E = K + Ugrav O
identify: We can’t use the constant-acceleration equations of
Chapter 2 because Throcky’s acceleration isn’t constant; the slope
decreases as he descends. Instead, we’ll use the energy approach.
Throcky moves along a circular arc, so we’ll also use what we
learned about circular motion in Section 5.4.
x
Solution
Solution
At y = h
At y = 0
ExamplE 7.4 spEEd at tHE bottom of a vErtiCal CirClE
Your cousin Throckmorton skateboards from rest down a curved,
frictionless ramp. If we treat Throcky and his skateboard as a particle, he moves through a quarter-circle with radius R = 3.00 m
(Fig. 7.9, next page). Throcky and his skateboard have a total mass
of 25.0 kg. (a) Find his speed at the bottom of the ramp. (b) Find
the normal force that acts on him at the bottom of the curve.
E = K + Ugrav
Set up: The only forces on Throcky are his weight and the norS
S
mal force n exerted by the ramp (Fig. 7.9b). Although n acts all
S
along the path, it does zero work because n is perpendicular to
Throcky’s displacement at every point. Hence Wother = 0 and
mechanical energy is conserved. We treat Throcky as a particle
located at the center of his body, take point 1 at the particle’s starting point, and take point 2 (which we let be y = 0) at the particle’s
low point. We take the positive y-direction upward; then y1 = R
and y2 = 0. Throcky starts at rest at the top, so v1 = 0. In part (a)
our target variable is his speed v2 at the bottom; in part (b) the target variable is the magnitude n of the normal force at point 2. To
find n, we’ll use Newton’s second law and the relation a = v 2>R.
Continued
210
Chapter 7 potential energy and energy Conservation
7.9 (a) Throcky skateboarding down a frictionless circular ramp. The total mechanical energy is constant.
(b) Free-body diagrams for Throcky and his skateboard at various points on the ramp.
(a)
Point 1 1y1 = R2
O
zero
v1 = 0
R = 3.00 m
E = K + Ugrav
At point 1
(b)
Point 1 n = 0
At each point, the normal force (n)
acts perpendicular to the direction w
of Throcky’s displacement, so only
n
the force of gravity (w) does work
on him.
Point 2 1y2 = 02
v2
R
n
w
zero
y = 0
w
Point 2
E = K + Ugrav
w
At point 2
w
exeCute: (a) The various energy quantities are
K1 = 0
K2 =
The y-component of Newton’s second law is
gFy = n + 1- w2 = marad = 2mg
Ugrav, 1 = mgR
1
2
2 mv 2
Ugrav, 2 = 0
n = w + 2mg = 3mg
From conservation of mechanical energy, Eq. (7.4),
K1 + Ugrav, 1 = K2 + Ugrav, 2
0 + mgR = 12mv 22 + 0
v2 = 22gR = 2219.80 m>s2213.00 m2 = 7.67 m>s
This answer doesn’t depend on the ramp being circular; Throcky
would have the same speed v2 = 12gR at the bottom of any ramp
of height R, no matter what its shape.
(b) To use Newton’s second law to find n at point 2, we need the
free-body diagram at that point (Fig. 7.9b). At point 2, Throcky is
moving at speed v2 = 12gR in a circle of radius R; his acceleration is toward the center of the circle and has magnitude
arad =
n
n
2gR
v 22
=
= 2g
R
R
= 3125.0 kg219.80 m>s22 = 735 N
At point 2 the normal force is three times Throcky’s weight. This
result doesn’t depend on the radius R of the ramp. We saw in
Examples 5.9 and 5.23 that the magnitude of n is the apparent
weight, so at the bottom of the curved part of the ramp Throcky
feels as though he weighs three times his true weight mg. But
when he reaches the horizontal part of the ramp, immediately to
the right of point 2, the normal force decreases to w = mg and
thereafter Throcky feels his true weight again. Can you see why?
evaluate: This example shows a general rule about the role of
forces in problems in which we use energy techniques: What matters is not simply whether a force acts, but whether that force
S
does work. If the force does no work, like the normal force n here,
then it does not appear in Eqs. (7.4) and (7.6).
Solution
ExamplE 7.5 a vErTiCal CirClE WiTH friCTion
Suppose that the ramp of Example 7.4 is not frictionless and that
Throcky’s speed at the bottom is only 6.00 m>s, not the 7.67 m>s
we found there. What work was done on him by the friction force?
solution
7.10 Energy bar graphs and free-body diagrams for Throcky
skateboarding down a ramp with friction.
identiFy and set uP: The setup is the same as in Example 7.4.
Figure 7.10 shows that again Sthe normal force does no work, but
now there is a friction force f that does do work Wf . Hence the
nongravitational work Wother done on Throcky between points 1
and 2 is equal to Wf and is not zero. Our target variable
is
S
Wf = Wother, which we’ll find by using Eq. (7.6). Since f points
opposite to Throcky’s motion, Wf is negative.
Point 1
The friction force
( f ) does negative work on
Throcky as he descends,
so the total mechanical
energy decreases.
f = 0
n = 0
w
f
n
w
R = 3.00 m
n
f
n
n
exeCute: The energy quantities are
K2 =
Ugrav, 2 = 0
=
1
2 125.0
kg216.00 m>s22 = 450 J
zero
1
2
2 mv 2
zero
K1 = 0
Ugrav, 1 = mgR = 125.0 kg219.80 m>s2213.00 m2 = 735 J
E = K + Ugrav
E = K + Ugrav
At point 1
At point 2
w
f
w
Point 2
f
w
7.1 Gravitational potential energy
From Eq. (7.6),
Wf = Wother
= K2 + Ugrav, 2 - K1 - Ugrav, 1
= 450 J + 0 - 0 - 735 J
= - 285 J
The work done by the friction force is -285 J, and the total
mechanical energy decreases by 285 J.
211
evaluate: Our result for Wf is negative. Can you see from the
free-body diagrams in Fig. 7.10 why this must be so?
It would be very difficult to apply Newton’s second law,
S
S
gF = ma, directly to this problem because the normal and
friction forces and the acceleration are continuously changing in
both magnitude and direction as Throcky descends. The energy
approach, by contrast, relates the motions at the top and bottom of
the ramp without involving the details of the motion in between.
Solution
ExamplE 7.6 an inClinEd planE WiTH friCTion
We want to slide a 12-kg crate up a 2.5-m-long ramp inclined at
30°. A worker, ignoring friction, calculates that he can do this by
giving it an initial speed of 5.0 m>s at the bottom and letting it
go. But friction is not negligible; the crate slides only 1.6 m up
the ramp, stops, and slides back down (Fig. 7.11a). (a) Find the
magnitude of the friction force acting on the crate, assuming that
it is constant. (b) How fast is the crate moving when it reaches the
bottom of the ramp?
done by friction as the crate slides back down, then use the energy
approach to find v3.
exeCute: (a) The energy quantities are
K1 = 12 112 kg215.0 m>s22 = 150 J
Ugrav, 1 = 0
K2 = 0
Ugrav, 2 = 112 kg219.8 m>s2210.80 m2 = 94 J
Wother = -fs
solution
identiFy and set uP: The friction force does work on the crate
as it slides from point 1, at the bottom of the ramp, to point 2,
where the crate stops instantaneously 1v2 = 02. Friction also does
work as the crate returns to the bottom of the ramp, which we’ll
call point 3 (Fig. 7.11a). We take the positive y-direction upward.
We take y = 0 (and hence Ugrav = 0) to be at ground level (point 1),
so y1 = 0, y2 = 11.6 m2sin 30° = 0.80 m, and y3 = 0. We are
given v1 = 5.0 m>s. In part (a) our target variable is f, the magnitude of the friction force as the crate slides up; we’ll find this by
using the energy approach. In part (b) our target variable is v3, the
crate’s speed at the bottom of the ramp. We’ll calculate the work
7.11 (a) A crate slides partway up the ramp, stops, and slides
back down. (b) Energy bar graphs for points 1, 2, and 3.
(a) The crate slides up from point 1
to point 2, then back down
to its starting position
m
2.5
(point 3).
m
1.6
Here s = 1.6 m. Using Eq. (7.6), we find
K1 + Ugrav, 1 + Wother = K2 + Ugrav, 2
Wother = - fs = 1K2 + Ugrav, 22 - 1K1 + Ugrav, 12
= 10 + 94 J2 - 1150 J + 02 = - 56 J = -fs
Wother
56 J
f =
=
= 35 N
s
1.6 m
The friction force of 35 N, acting over 1.6 m, causes the mechanical energy of the crate to decrease from 150 J to 94 J (Fig. 7.11b).
(b) As the crate moves from point 2 to point 3, the work done
by friction has the same negative value as from point 1 to point 2.
(Both the friction force and the displacement reverse direction,
but their magnitudes don’t change.) The total work done by friction between points 1 and 3 is therefore
Wother = Wfric = -2fs = -2156 J2 = -112 J
Point
2
From part (a), K1 = 150 J and Ugrav, 1 = 0; in addition, Ugrav, 3 = 0
since y3 = 0. Equation (7.6) then gives
v2 = 0
K1 + Ugrav, 1 + Wother = K3 + Ugrav, 3
K3 = K1 + Ugrav, 1 - Ugrav, 3 + Wother
/
v1 = 5.0 m s
The crate is moving
at speed v3 when it
returns to point 3.
30°
= 150 J + 0 - 0 + 1- 112 J2 = 38 J
0.80 m
The crate returns to the bottom of the ramp with only 38 J of the
original 150 J of mechanical energy (Fig. 7.11b). Since K3 = 12 mv 32,
Point 1 , 3
v3 =
The force of friction does negative work on
the crate as it moves, so the total mechanical
energy E = K + Ugrav decreases.
(b)
E = K + Ugrav
At point 2
evaluate: Energy is lost due to friction, so the crate’s speed
zero
At point 1
zero
zero
E = K + Ugrav
2138 J2
2K3
= 2.5 m>s
=
B m
B 12 kg
E = K + Ugrav
At point 3
v3 = 2.5 m>s when it returns to the bottom of the ramp is less
than the speed v1 = 5.0 m>s at which it left that point. In part (b)
we applied Eq. (7.6) to points 1 and 3, considering the round trip
as a whole. Alternatively, we could have considered the second
part of the motion by itself and applied Eq. (7.6) to points 2 and 3.
Try it; do you get the same result for v3?
212
Chapter 7 potential energy and energy Conservation
TesT Your undersTanding of secTion 7.1 The figure shows two frictionless ramps. The heights y1 and y2 are the same for both ramps. If a block of mass m is
released from rest at the left-hand end of each ramp, which block arrives at the right-hand
end with the greater speed? (i) Block I; (ii) block II; (iii) the speed is the same for both
blocks.
y1
Block I
m
Block II
m
y1
y2
y2
❙
7.2 elasTic PoTenTial energY
7.12 The Achilles tendon, which runs
along the back of the ankle to the heel
bone, acts like a natural spring. When it
stretches and then relaxes, this tendon
stores and then releases elastic potential
energy. This spring action reduces the
amount of work your leg muscles must do
as you run.
Achilles
tendon
Calf muscle
Heel bone
In many situations we encounter potential energy that is not gravitational in
nature. One example is a rubber-band slingshot. Work is done on the rubber
band by the force that stretches it, and that work is stored in the rubber band
until you let it go. Then the rubber band gives kinetic energy to the projectile.
This is the same pattern we saw with the baseball in Example 7.2: Do work on
the system to store energy, which can later be converted to kinetic energy. We’ll
describe the process of storing energy in a deformable body such as a spring or
rubber band in terms of elastic potential energy (Fig. 7.12). A body is called
elastic if it returns to its original shape and size after being deformed.
To be specific, we’ll consider storing energy in an ideal spring, like the ones we
discussed in Section 6.3. To keep such an ideal spring stretched by a distance x,
Many
we must exert a force F = kx, where k is the force constant of the spring.
S
elastic bodies show this same direct proportionality between force F and displacement x, provided that x is sufficiently small.
Let’s proceed just as we did for gravitational potential energy. We begin with
the work done by the elastic (spring) force and then combine this with the work–
energy theorem. The difference is that gravitational potential energy is a shared
property of a body and the earth, but elastic potential energy is stored in just the
spring (or other deformable body).
Figure 7.13 shows the ideal spring from Fig. 6.18 but with its left end held stationary and its right end attached to a block with mass m that can move along the
x-axis. In Fig. 7.13a the block is at x = 0 when the spring is neither stretched nor
compressed. We move the block to one side, thereby stretching or compressing
the spring, then let it go. As the block moves from a different position x1 to a different position x2 , how much work does the elastic (spring) force do on the block?
We found in Section 6.3 that the work we must do on the spring to move one
end from an elongation x1 to a different elongation x2 is
W = 12 kx 22 - 12 kx 12
(work done on a spring)
(7.8)
where k is the force constant of the spring. If we stretch the spring farther, we do
positive work on the spring; if we let the spring relax while holding one end, we
do negative work on it. This expression for work is also correct when the spring
is compressed such that x1, x2, or both are negative. Now, from Newton’s third
law the work done by the spring is just the negative of the work done on the
spring. So by changing the signs in Eq. (7.8), we find that in a displacement from
x1 to x2 the spring does an amount of work Wel given by
Wel = 12 kx 12 - 12 kx 22
(work done by a spring)
(7.9)
The subscript “el” stands for elastic. When both x1 and x2 are positive and
x2 7 x1 (Fig. 7.13b), the spring does negative work on the block, which moves
in the +x@direction while the spring pulls on it in the -x@direction. The spring
stretches farther, and the block slows down. When both x1 and x2 are positive
213
7.2 elastic potential energy
and x2 6 x1 (Fig. 7.13c), the spring does positive work as it relaxes and the block
speeds up. If the spring can be compressed as well as stretched, x1, x2, or both
may be negative, but the expression for Wel is still valid. In Fig. 7.13d, both x1 and
x2 are negative, but x2 is less negative than x1 ; the compressed spring does positive work as it relaxes, speeding the block up.
Just as for gravitational work, we can express Eq. (7.9) for the work done by
the spring in terms of a quantity at the beginning and end of the displacement.
This quantity is 12 kx 2, and we define it to be the elastic potential energy:
7.13 Calculating the work done by a
spring attached to a block on a horizontal
surface. The quantity x is the extension or
compression of the spring.
(a)
Here the spring is
x = 0 neither stretched
m
nor compressed.
O
Force constant of spring
Elastic potential energy
stored in a spring
Uel = 12 kx 2
Elongation of spring
(x 7 0 if stretched,
x 6 0 if compressed)
(7.10)
(b)
Figure 7.14 is a graph of Eq. (7.10). As for all other energy and work quantities,
the unit of Uel is the joule (J); to see this from Eq. (7.10), recall that the units of
k are N>m and that 1 N # m = 1 J. We can now use Eq. (7.10) to rewrite Eq. (7.9)
for the work Wel done by the spring:
Work done by the elastic force ...
Wel =
1
1
2
2
2 kx1 - 2 kx2
Force constant of spring
... equals the negative of the
change in elastic potential energy.
= Uel, 1
-
Uel, 2 =
-∆Uel
As the spring stretches, it does negative
S
work on the block.
s
x2
x1
m
O
(c)
Initial and final elongations of spring
As the spring relaxes, it does positive
work on the block. Ss
x1
x2
m
O
x
S
Fspring
(d)
S
s
x1
A compressed spring
also does positive
work on the block as
it relaxes.
x2
Caution gravitational potential energy vs. elastic potential energy An important dif-
ference between gravitational potential energy Ugrav = mgy and elastic potential energy
Uel = 12 kx 2 is that we cannot choose x = 0 to be wherever we wish. In Eq. (7.10), x = 0
must be the position at which the spring is neither stretched nor compressed. At that position, both its elastic potential energy and the force that it exerts are zero. ❙
x
S
Fspring
(7.11)
When a stretched spring is stretched farther, as in Fig. 7.13b, Wel is negative
and Uel increases; more elastic potential energy is stored in the spring. When
a stretched spring relaxes, as in Fig. 7.13c, x decreases, Wel is positive, and Uel
decreases; the spring loses elastic potential energy. Figure 7.14 shows that Uel is
positive for both positive and negative x values; Eqs. (7.10) and (7.11) are valid for
both cases. The more a spring is compressed or stretched, the greater its elastic
potential energy.
m
S
x
O
Fspring
The work–energy theorem says that Wtot = K2 - K1 , no matter what kind of
forces are acting on a body. If the elastic force is the only force that does work on
the body, then
Wtot = Wel = Uel, 1 - Uel, 2
and so
7.14 The graph of elastic potential
energy for an ideal spring is a parabola:
Uel = 12 kx 2, where x is the extension or
compression of the spring. Elastic potential energy Uel is never negative.
If only the elastic force does work, total mechanical energy is conserved:
Initial kinetic energy
Initial elastic potential energy
K1 = 12 mv12
Uel, 1 = 12 kx 12
K1 + Uel,1 = K2 + Uel, 2
Final kinetic energy
K2 = 12 mv22
Uel
(7.12)
Final elastic potential energy
Uel, 2 = 12 kx 22
In this case the total mechanical energy E = K + Uel—the sum of kinetic and
elastic potential energies—is conserved. An example of this is the motion of the
block in Fig. 7.13, provided the horizontal surface is frictionless so no force does
work other than that exerted by the spring.
For Eq. (7.12) to be strictly correct, the ideal spring that we’ve been discussing
must also be massless. If the spring has mass, it also has kinetic energy as the
x
Spring is
compressed:
x 6 0.
O
Spring is
stretched:
x 7 0.
x
214
Chapter 7 potential energy and energy Conservation
BIO application elastic Potential
energy of a Cheetah When a cheetah
(Acinonyx jubatus) gallops, its back flexes and
extends dramatically. Flexion of the back
stretches tendons and muscles along the top
of the spine and also compresses the spine,
storing elastic potential energy. When the
cheetah launches into its next bound, this
energy is released, enabling the cheetah to
run more efficiently.
coils of the spring move back and forth. We can ignore the kinetic energy of the
spring if its mass is much less than the mass m of the body attached to the spring.
For instance, a typical automobile has a mass of 1200 kg or more. The springs in
its suspension have masses of only a few kilograms, so their mass can be ignored
if we want to study how a car bounces on its suspension.
situations with Both Gravitational and
elastic Potential energy
Equation (7.12) is valid when the only potential energy in the system is elastic potential energy. What happens when we have both gravitational and elastic
forces, such as a block attached to the lower end of a vertically hanging spring?
And what if work is also done by other forces that cannot be described in terms
of potential energy, such as the force of air resistance on a moving block? Then
the total work is the sum of the work done by the gravitational force 1Wgrav2, the
work done by the elastic force 1Wel2, and the work done by other forces 1Wother2:
Wtot = Wgrav + Wel + Wother . The work–energy theorem then gives
Wgrav + Wel + Wother = K2 - K1
The work done by the gravitational force is Wgrav = Ugrav, 1 - Ugrav, 2 and the
work done by the spring is Wel = Uel, 1 - Uel, 2 . Hence we can rewrite the work–
energy theorem for this most general case as
(valid in
K1 + Ugrav, 1 + Uel, 1 + Wother = K2 + Ugrav, 2 + Uel, 2
general) (7.13)
or, equivalently,
General relationship for kinetic energy and potential energy:
Initial kinetic energy
Final kinetic energy
Difference in nose-to-tail length
7.15 Trampoline jumping involves an
interplay among kinetic energy, gravitational potential energy, and elastic potential energy. Due to air resistance and
friction forces within the trampoline,
mechanical energy is not conserved.
That’s why the bouncing eventually stops
unless the jumper does work with his or
her legs to compensate for the lost energy.
Gravitational potential energy
increases as jumper ascends.
Kinetic
energy
increases as
jumper
moves faster.
Elastic potential energy increases
when trampoline is stretched.
K1 + U1 + Wother = K2 + U2
Initial potential
energy of all kinds
Work done by other forces
(not associated with potential energy)
(7.14)
Final potential
energy of all kinds
where U = Ugrav + Uel = mgy + 12 kx 2 is the sum of gravitational potential
energy and elastic potential energy. We call U simply “the potential energy.”
Equation (7.14) is the most general statement of the relationship among kinetic
energy, potential energy, and work done by other forces. It says:
The work done by all forces other than the gravitational force or elastic force
equals the change in the total mechanical energy E = K + U of the system.
The “system” is made up of the body of mass m, the earth with which it interacts
through the gravitational force, and the spring of force constant k.
If Wother is positive, E = K + U increases; if Wother is negative, E decreases.
If the gravitational and elastic forces are the only forces that do work on the
body, then Wother = 0 and the total mechanical energy E = K + U is conserved.
[Compare Eq. (7.14) to Eqs. (7.6) and (7.7), which include gravitational potential
energy but not elastic potential energy.]
Trampoline jumping (Fig. 7.15) involves transformations among kinetic
energy, elastic potential energy, and gravitational potential energy. As the jumper
descends through the air from the high point of the bounce, gravitational
potential energy Ugrav decreases and kinetic energy K increases. Once the jumper
touches the trampoline, some of the mechanical energy goes into elastic potential
energy Uel stored in the trampoline’s springs. At the lowest point of the trajectory (Ugrav is minimum), the jumper comes to a momentary halt 1K = 02 and
the springs are maximally stretched 1Uel is maximum2. The springs then convert
their energy back into K and Ugrav, propelling the jumper upward.
7.2 elastic potential energy
problEm-solving sTraTEgy 7.2
215
problEms Using mECHaniCal EnErgy ii
Problem-Solving Strategy 7.1 (Section 7.1) is useful in solving problems that involve elastic forces as well as gravitational
forces. The only new wrinkle is that the potential energy U now
includes the elastic potential energy Uel = 12 kx 2, where x is the
displacement of the spring from its unstretched length. The work
done by the gravitational and elastic forces is accounted for by
their potential energies; the work done by other forces, Wother,
must still be included separately.
Solution
ExamplE 7.7 moTion WiTH ElasTiC poTEnTial EnErgy
A glider with mass m = 0.200 kg sits on a frictionless, horizontal
air track, connected to a spring with force constant k = 5.00 N>m.
You pull on the glider, stretching the spring 0.100 m, and release
it from rest. The glider moves back toward its equilibrium position
1x = 02. What is its x-velocity when x = 0.080 m?
solution
identiFy and set uP: As the glider starts to move, elastic
potential energy is converted to kinetic energy. The glider remains
at the same height throughout the motion, so gravitational potential energy is not a factor and U = Uel = 12 kx 2. Figure 7.16 shows
our sketches. Only the spring force does work on the glider, so
Wother = 0 in Eq. (7.14). We designate the point where the glider
is released as point 1 1that is, x1 = 0.100 m2 and x2 = 0.080 m as
point 2. We are given v1x = 0; our target variable is v2x .
exeCute: The energy quantities are
K1 = 12 mv 1x2 = 1210.200 kg21022 = 0
U1 = 12 kx 12 = 1215.00 N>m210.100 m22 = 0.0250 J
K2 = 12 mv 2x2
U2 = 12 kx 22 = 1215.00 N>m210.080 m22 = 0.0160 J
We use Eq. (7.14) with Wother = 0 to solve for K2 and then find v2x :
K2 = K1 + U1 - U2 = 0 + 0.0250 J - 0.0160 J = 0.0090 J
7.16 Our sketches and energy bar graphs for this problem.
v2x = {
210.0090 J2
2K2
= {0.30 m>s
= {
B m
B 0.200 kg
We choose the negative root because the glider is moving in the
-x@direction. Our answer is v2x = -0.30 m>s.
evaluate: Eventually the spring will reverse the glider’s motion,
pushing it back in the + x-direction (see Fig. 7.13d). The solution
v2x = +0.30 m>s tells us that when the glider passes through
x = 0.080 m on this return trip, its speed will be 0.30 m>s, just as
when it passed through this point while moving to the left.
Solution
ExamplE 7.8
moTion WiTH ElasTiC poTEnTial EnErgy and Work donE by oTHEr forCEs
Suppose the glider in Example 7.7 is initially at rest at x = 0, with
the spring unstretched.
You then push on the glider with a conS
stant force F 1magnitude 0.610 N2 in the + x-direction. What is
the glider’s velocity when it has moved to x = 0.100 m?
solution
S
identiFy and set uP: Although the force F you apply is con-
stant, the spring force isn’t, so the acceleration of the glider won’t
be constant. Total mechanical
energy is not conserved because
S
of the work done by force F, so Wother in Eq. (7.14) is not zero.
As in Example 7.7, we ignore gravitational potential energy
because the glider’s height doesn’t change. Hence we again have
U = Uel = 12 kx 2. This time, we let point 1 be at x1 = 0, where the
velocity is v1x = 0, and let point 2 be at x = 0.100 m. The glider’s
displacement is then ∆x = x2 - x1 = 0.100 m. Our target variable is v2x, the velocity at point 2.
S
exeCute: Force F is constant and in the same direction as the
displacement, so the work done by this force is F∆x. Then the
energy quantities are
K1 = 0
U1 = 12 kx 12 = 0
K2 = 12 mv 2x 2
U2 = 12 kx 22 = 12 15.00 N>m210.100 m22 = 0.0250 J
Wother = F∆x = 10.610 N210.100 m2 = 0.0610 J
Continued
216
Chapter 7 potential energy and energy Conservation
The
initial total mechanical energy is zero; the work done by
S
F increases the total mechanical energy to 0.0610 J, of which
U2 = 0.0250 J is elastic potential energy. The remainder is kinetic
energy. From Eq. (7.14),
K1 + U1 + Wother = K2 + U2
K2 = K1 + U1 + Wother - U2
= 0 + 0 + 0.0610 J - 0.0250 J = 0.0360 J
v2x =
2K2
B m
=
210.0360 J2
B 0.200 kg
= 0.60 m>s
We choose the positive square root because the glider is moving in
the + x-direction.
evaluate: What would be different if we disconnected the glider
S
from the spring? Then only F would do work, there would be
zero elastic potential energy at all times, and Eq. (7.14) would
give us
K2 = K1 + Wother = 0 + 0.0610 J
v2x =
210.0610 J2
2K2
= 0.78 m>s
=
B m
B 0.200 kg
Our answer v2x = 0.60 m>s is less than 0.78 m>s because the spring
does negative work on the glider as it stretches (see Fig. 7.13b).
If you stop pushing on the glider when it reaches x = 0.100 m,
only the spring force does work on it thereafter. Hence for
x 7 0.100 m, the total mechanical energy E = K + U = 0.0610 J
is constant. As the spring continues to stretch, the glider slows
down and the kinetic energy K decreases as the potential
energy increases. The glider comes to rest at some point x = x3,
at which the kinetic energy is zero and the potential energy
U = Uel = 12 kx 32 equals the total mechanical energy 0.0610 J.
Can you show that x3 = 0.156 m? (It moves an additional 0.056 m
after you stop pushing.) If there is no friction, will the glider remain
at rest?
Solution
ExamplE 7.9 motion witH gravitational, ElastiC, and friCtion forCEs
A 2000-kg 119,600@N2 elevator with broken cables in a test rig
is falling at 4.00 m>s when it contacts a cushioning spring at the
bottom of the shaft. The spring is intended to stop the elevator,
compressing 2.00 m as it does so (Fig. 7.17). During the motion
a safety clamp applies a constant 17,000-N friction force to the
elevator. What is the necessary force constant k for the spring?
Solution
identify and Set up: We’ll use the energy approach and
Eq. (7.14) to determine k, which appears in the expression for
elastic potential energy. This problem involves both gravitational
and elastic potential energies. Total mechanical energy is not conserved because the friction force does negative work Wother on
the elevator. We take point 1 as the position of the bottom of the
elevator when it contacts the spring, and point 2 as its position
when it stops. We choose the origin to be at point 1, so y1 = 0
7.17 The fall of an elevator is stopped by a spring and by a
constant friction force.
and y2 = -2.00 m. With this choice the coordinate of the upper
end of the spring after contact is the same as the coordinate of
the elevator, so the elastic potential energy at any point between
points 1 and 2 is Uel = 12 ky2. The gravitational potential energy
is Ugrav = mgy as usual. We know the initial and final speeds of
the elevator and the magnitude of the friction force, so the only
unknown is the force constant k (our target variable).
exeCute: The elevator’s initial speed is v1 = 4.00 m>s, so its initial kinetic energy is
K1 = 12 mv 12 = 12 12000 kg214.00 m>s22 = 16,000 J
The elevator stops at point 2, so K2 = 0. At point 1 the potential
energy U1 = Ugrav + Uel is zero; Ugrav is zero because y1 = 0,
and Uel = 0 because the spring is uncompressed. At point 2 there
are both gravitational and elastic potential energies, so
U2 = mgy2 + 12 ky22
The gravitational potential energy at point 2 is
mgy2 = 12000 kg219.80 m>s221-2.00 m2 = - 39,200 J
f = 17,000 N
The “other” force is the constant 17,000-N friction force. It acts
opposite to the 2.00-m displacement, so
Wother = -117,000 N212.00 m2 = -34,000 J
We put these terms into Eq. (7.14), K1 + U1 + Wother = K2 + U2:
v1 =
4.00 m>s
m =
2000 kg
v2 = 0
Point 1
w= mg
K1 + 0 + Wother = 0 + 1mgy2 + 12 ky222
k =
21K1 + Wother - mgy22
y22
2316,000 J + 1- 34,000 J2 - 1- 39,200 J24
2.00 m
=
Point 2
= 1.06 * 104 N>m
1-2.00 m22
This is about one-tenth the force constant of a spring in an automobile suspension.
7.3 Conservative and Nonconservative Forces
evaluate: There might seem to be a paradox here. The elastic
potential energy at point 2 is
1
2
2 ky2
= 12 11.06 * 104 N>m21- 2.00 m22 = 21,200 J
This is more than the total mechanical energy at point 1:
E 1 = K1 + U1 = 16,000 J + 0 = 16,000 J
But the friction force decreased the mechanical energy of the system by 34,000 J between points 1 and 2. Did energy appear from
nowhere? No. At point 2, which is below the origin, there is also
negative gravitational potential energy mgy2 = - 39,200 J. The
total mechanical energy at point 2 is therefore not 21,200 J but
217
This is just the initial mechanical energy of 16,000 J minus 34,000 J
lost to friction.
Will the elevator stay at the bottom of the shaft? At point 2 the
compressed spring exerts an upward force of magnitude Fspring =
11.06 * 104 N>m212.00 m2 = 21,200 N, while the downward
force of gravity is only w = mg = 12000 kg219.80 m>s22 =
19,600 N. If there were no friction, there would be a net upward
force of 21,200 N - 19,600 N = 1600 N, and the elevator would
rebound. But the safety clamp can exert a kinetic friction force of
17,000 N, and it can presumably exert a maximum static friction
force greater than that. Hence the clamp will keep the elevator
from rebounding.
E 2 = K2 + U2 = 0 + 12 ky22 + mgy2
= 0 + 21,200 J + 1-39,200 J2 = -18,000 J
test your understandinG oF seCtion 7.2 Consider the situation in
Example 7.9 at the instant when the elevator is still moving downward and the spring
is compressed by 1.00 m. Which of the energy bar graphs in the figure most accurately
shows the kinetic energy K, gravitational potential energy Ugrav , and elastic potential
energy Uel at this instant?
(i)
( ii)
(iii)
K Ugrav
K Ugrav Uel
Ugrav
Uel
K
data SpeakS
(iv)
Conservation of energy
Uel
Uel
K Ugrav
❙
7.3 Conservative and
nonConservative ForCes
In our discussions of potential energy we have talked about “storing” kinetic
energy by converting it to potential energy, with the idea that we can retrieve it
again as kinetic energy. For example, when you throw a ball up in the air, it slows
down as kinetic energy is converted to gravitational potential energy. But on the
way down the ball speeds up as potential energy is converted back to kinetic
energy. If there is no air resistance, the ball is moving just as fast when you catch
it as when you threw it.
Another example is a glider moving on a frictionless horizontal air track that
runs into a spring bumper. The glider compresses the spring and then bounces
back. If there is no friction, the glider ends up with the same speed and kinetic
energy it had before the collision. Again, there is a two-way conversion from
kinetic to potential energy and back. In both cases the total mechanical energy,
kinetic plus potential, is constant or conserved during the motion.
Conservative Forces
A force that offers this opportunity of two-way conversion between kinetic and
potential energies is called a conservative force. We have seen two examples of
conservative forces: the gravitational force and the spring force. (Later in this
book we’ll study another conservative force, the electric force between charged
objects.) An essential feature of conservative forces is that their work is always
reversible. Anything that we deposit in the energy “bank” can later be withdrawn
without loss. Another important aspect of conservative forces is that if a body
follows different paths from point 1 to point 2, the work done by a conservative
force is the same for all of these paths (Fig. 7.18). For example, if a body stays
S
close to the surface of the earth, the gravitational force mg is independent of
When students were given a problem
about conservation of mechanical energy
for motion along a curved path, more
than 32% gave an incorrect answer.
Common errors:
●
●
Forgetting that the change in gravitational potential energy along a curved
path depends on only the difference
between the final and initial heights, not
the shape of the path.
Forgetting that if gravity is the only force
that does work, mechanical energy is
conserved. Then the change in kinetic
energy along the path is determined
by solely the change in gravitational
potential energy. The shape of the path
doesn’t matter.
7.18 The work done by a conservative
force such as gravity depends on only the
endpoints of a path, not the specific path
taken between those points.
Because the gravitational force is conservative,
the work it does is the same for all three paths.
Final
position
Initial
position
218
Chapter 7 potential energy and energy Conservation
PhET: The Ramp
height, and the work done by this force depends on only the change in height. If
the body moves around a closed path, ending at the same height where it started,
the total work done by the gravitational force is always zero.
In summary, the work done by a conservative force has four properties:
1. It can be expressed as the difference between the initial and final values of
a potential-energy function.
2. It is reversible.
3. It is independent of the path of the body and depends on only the starting
and ending points.
4. When the starting and ending points are the same, the total work is zero.
When the only forces that do work are conservative forces, the total mechanical
energy E = K + U is constant.
nonconservative Forces
Not all forces are conservative. Consider the friction force acting on the crate
sliding on a ramp in Example 7.6 (Section 7.1). When the body slides up and then
back down to the starting point, the total work done on it by the friction force is
not zero. When the direction of motion reverses, so does the friction force, and
friction does negative work in both directions. Friction also acts when a car with
its brakes locked skids with decreasing speed (and decreasing kinetic energy).
The lost kinetic energy can’t be recovered by reversing the motion or in any other
way, and mechanical energy is not conserved. So there is no potential-energy
function for the friction force.
In the same way, the force of fluid resistance (see Section 5.3) is not conservative. If you throw a ball up in the air, air resistance does negative work on the ball
while it’s rising and while it’s descending. The ball returns to your hand with less
speed and less kinetic energy than when it left, and there is no way to get back
the lost mechanical energy.
A force that is not conservative is called a nonconservative force. The work
done by a nonconservative force cannot be represented by a potential-energy
function. Some nonconservative forces, like kinetic friction or fluid resistance,
cause mechanical energy to be lost or dissipated; a force of this kind is called a
dissipative force. There are also nonconservative forces that increase mechanical energy. The fragments of an exploding firecracker fly off with very large
kinetic energy, thanks to a chemical reaction of gunpowder with oxygen. The
forces unleashed by this reaction are nonconservative because the process is
not reversible. (The fragments never spontaneously reassemble themselves into
a complete firecracker!)
friCTional Work dEpEnds on THE paTH
You are rearranging your furniture and wish to move a 40.0-kg
futon 2.50 m across the room. A heavy coffee table, which you
don’t want to move, blocks this straight-line path. Instead, you
slide the futon along a dogleg path; the doglegs are 2.00 m and
1.50 m long. How much more work must you do to push the futon
along the dogleg path than along the straight-line path? The coefficient of kinetic friction is mk = 0.200.
solution
identiFy and set uP: Here both you and friction do work on the
futon, so we must use the energy relationship that includes “other”
forces. We’ll use this relationship to find a connection between
the work that you do and the work that friction does. Figure 7.19
Solution
ExamplE 7.10
shows our sketch. The futon is at rest at both point 1 and point 2,
so K1 = K2 = 0. There is no elastic potential energy (there
are no springs), and the gravitational potential energy does not
7.19 Our sketch for this problem.
7.3 Conservative and Nonconservative Forces
change because the futon moves only horizontally, so U1 = U2.
From Eq. (7.14) it follows that Wother = 0. That “other” work done
on the futon is the sum of the positive work you do, Wyou, and the
negative work done by friction, Wfric. Since the sum of these is
zero, we have
219
Wyou = -Wfric = -1-fks2 = + mkmgs
= 10.2002140.0 kg219.80 m>s2212.50 m2
= 196 J
(straight-line path)
Wyou = -Wfric = + mkmgs
Wyou = - Wfric
= 10.2002140.0 kg219.80 m>s2212.00 m + 1.50 m2
= 274 J
(dogleg path)
So we can calculate the work done by friction to determine
Wyou .
The extra work you must do is 274 J - 196 J = 78 J.
exeCute: The floor is horizontal, so the normal force on the
evaluate: Friction does different amounts of work on the futon,
futon equals its weight mg and the magnitude of the friction force
is fk = mkn = mkmg. The work you do over each path is then
-196 J and -274 J, on these different paths between points 1
and 2. Hence friction is a nonconservative force.
Solution
ExamplE 7.11
ConsErvaTivE or nonConsErvaTivE?
S
In a region of space the force on an electron is F = Cxen, where C
is a positive constant. The electron moves around a square loop in
the xy-plane
(Fig. 7.20). Calculate the work done on the electron
S
by force F during a counterclockwise trip around the square. Is
this force conservative or nonconservative?
and the workS done on the first leg is W1 = 0. The force has the
same value F = CLen everywhere on the second leg, from 1L, 02
toS 1L, L2. The displacement on this leg is in the + y@direction, so
dl = dyen and
S
S
F ~ dl = CLen ~ dyen = CL dy
solution
S
identiFy and set uP: Force F is not constant and in general is
not in the same
direction as the displacement. To calculate the
S
work done by F, we’ll use the general expression Eq. (6.14):
P2
W =
LP1
S
S
F ~ dl
S
where dl is an infinitesimal displacement. We’ll calculate the
work done on each leg of the square separately, and add the results
to find the work
done on the round trip. If this round-trip work is
S
zero, force F is conservative and can be represented by a potentialenergy function.
exeCute: On the first leg, from 10, 02 to 1L, 02, the force is
S
S
everywhere perpendicular to the displacement. So F ~ dl = 0,
7.20 An electron moving
around a square loop while being
S
acted on by the force F = Cxen.
S
F
y
(0, L)
S
F=0
(L, L)
Leg 3
S
dl
S
S
dl
dl
Leg 4
Leg 2
S
F = CLen
S
F
(0, 0)
Leg 1
S
dl
(L, 0)
x
The work done on the second leg is then
W2 =
1L, L2
L1L, 02
S
S
F ~ dl =
y=L
Ly = 0
CL dy = CL
L0
L
dy = CL2
S
On the third leg, from 1L, L2 to 10, L2, F is again perpendicular
to the displacement and so W3 = 0. The force is zero on the
final
S
leg, from 10, L2 to 10, 02, so W4 = 0. The work done by F on the
round trip is therefore
W = W1 + W2 + W3 + W4 = 0 + CL2 + 0 + 0 = CL2
The starting
and ending points are the same, but the total work
S
done by F is not zero. This is a nonconservative force; it cannot
be represented by a potential-energy function.
evaluate: Because W 7 0, the mechanical energy increases as
the electron goes around the loop. This is actually what happens
in an electric generating plant: A loop of wire is moved through a
magnetic field, which gives rise to a nonconservative force similar
to the one here. Electrons in the wire gain energy as they move
around the loop, and this energy is carried via transmission lines
to the consumer. (We’ll discuss this in Chapter 29.) S
If the electron went clockwise around the loop, F would be un-S
affected but the direction of each infinitesimal displacement dl
would be reversed. Thus the sign of work would also reverse,
and the work for a clockwise round trip would be W = - CL2.
This is a different behavior than the nonconservative friction
force. The work done by friction on a body that slides in any direction over a stationary surface is always negative (see Example 7.6
in Section 7.1).
220
Chapter 7 potential energy and energy Conservation
application nonconservative
Forces and internal energy in a
tire An automobile tire deforms and flexes
like a spring as it rolls, but it is not an ideal
spring: Nonconservative internal friction forces
act within the rubber. As a result, mechanical
energy is lost and converted to internal energy
of the tire. Thus the temperature of a tire
increases as it rolls, which causes the pressure
of the air inside the tire to increase as well.
That’s why tire pressures are best checked
before the car is driven, when the tire is cold.
the law of Conservation of energy
Nonconservative forces cannot be represented in terms of potential energy. But
we can describe the effects of these forces in terms of kinds of energy other than
kinetic or potential energy. When a car with locked brakes skids to a stop, both
the tires and the road surface become hotter. The energy associated with this
change in the state of the materials is called internal energy. Raising the temperature of a body increases its internal energy; lowering the body’s temperature
decreases its internal energy.
To see the significance of internal energy, let’s consider a block sliding on
a rough surface. Friction does negative work on the block as it slides, and the
change in internal energy of the block and surface (both of which get hotter)
is positive. Careful experiments show that the increase in the internal energy is
exactly equal to the absolute value of the work done by friction. In other words,
∆Uint = -Wother
where ∆Uint is the change in internal energy. We substitute this into Eq. (7.14):
K1 + U1 - ∆Uint = K2 + U2
Writing ∆K = K2 - K1 and ∆U = U2 - U1 , we can finally express this as
Law of conservation of energy:
∆K + ∆U + ∆Uint = 0
Change in kinetic energy
7.21 The battery pack in this radiocontrolled helicopter contains 2.4 * 104 J
of electric energy. When this energy is
used up, the internal energy of the
battery pack decreases by this amount,
so ∆Uint = -2.4 * 104 J. This energy
can be converted to kinetic energy to
make the rotor blades and helicopter go
faster, or to gravitational potential energy
to make the helicopter climb.
Change in internal energy
This remarkable statement is the general form of the law of conservation of
energy. In a given process, the kinetic energy, potential energy, and internal
energy of a system may all change. But the sum of those changes is always zero.
If there is a decrease in one form of energy, it is made up for by an increase in the
other forms (Fig. 7.21). When we expand our definition of energy to include
internal energy, Eq. (7.15) says: Energy is never created or destroyed; it only
changes form. No exception to this rule has ever been found.
The concept of work has been banished from Eq. (7.15); instead, it suggests that we think purely in terms of the conversion of energy from one
form to another. For example, when you throw a baseball straight up, you convert a portion of the internal energy of your molecules to kinetic energy of the
baseball. This is converted to gravitational potential energy as the ball climbs
and back to kinetic energy as the ball falls. If there is air resistance, part of the
energy is used to heat up the air and the ball and increase their internal energy.
Energy is converted back to the kinetic form as the ball falls. If you catch the ball
in your hand, whatever energy was not lost to the air once again becomes internal
energy; the ball and your hand are now warmer than they were at the beginning.
In Chapters 19 and 20, we will study the relationship of internal energy to
temperature changes, heat, and work. This is the heart of the area of physics
called thermodynamics.
?
Solution
ConCEpTUal ExamplE 7.12
Change in potential energy
(7.15)
Work donE by friCTion
Let’s return to Example 7.5 (Section 7.1), in which Throcky
skateboards down a curved ramp. He starts with zero kinetic
energy and 735 J of potential energy, and at the bottom he
has 450 J of kinetic energy and zero potential energy; hence
∆K = +450 J and ∆U = - 735 J. The work Wother = Wfric done
by the friction forces is - 285 J, so the change in internal energy is
∆Uint = -Wother = + 285 J. The skateboard wheels and bearings
and the ramp all get a little warmer. In accordance with Eq. (7.15),
the sum of the energy changes equals zero:
∆K + ∆U + ∆Uint = +450 J + 1- 735 J2 + 285 J = 0
The total energy of the system (including internal, nonmechanical
forms of energy) is conserved.
7.4 Force and potential energy
TesT Your undersTanding of secTion 7.3 In a hydroelectric generating
station, falling water is used to drive turbines (“water wheels”), which in turn run electric generators. Compared to the amount of gravitational potential energy released by the
falling water, how much electrical energy is produced? (i) The same; (ii) more; (iii) less. ❙
7.4 force and PoTenTial energY
For the two kinds of conservative forces (gravitational and elastic) we have
studied, we started with a description of the behavior of the force and derived
from that an expression for the potential energy. For example, for a body with
mass m in a uniform gravitational field, the gravitational force is Fy = -mg.
We found that the corresponding potential energy is U1y2 = mgy. The force
that an ideal spring exerts on a body is Fx = -kx, and the corresponding
potential-energy function is U1x2 = 12 kx 2.
In studying physics, however, you’ll encounter situations in which you are
given an expression for the potential energy as a function of position and have to
find the corresponding force. We’ll see several examples of this kind when we
study electric forces later in this book: It’s often far easier to calculate the electric potential energy first and then determine the corresponding electric force
afterward.
Here’s how we find the force that corresponds to a given potential-energy
expression. First let’s consider motion along a straight line, with coordinate x.
We denote the x-component of force, a function of x, by Fx 1x2 and the potential
energy as U1x2. This notation reminds us that both Fx and U are functions of x.
Now we recall that in any displacement, the work W done by a conservative force
equals the negative of the change ∆U in potential energy:
W = - ∆U
Let’s apply this to a small displacement ∆x. The work done by the force Fx 1x2
during this displacement is approximately equal to Fx 1x2 ∆x. We have to say
“approximately” because Fx 1x2 may vary a little over the interval ∆x. So
∆U
∆x
You can probably see what’s coming. We take the limit as ∆x S 0; in this limit,
the variation of Fx becomes negligible, and we have the exact relationship
Fx 1x2 ∆x = - ∆U
Force from potential energy:
In one-dimensional motion,
the value of a conservative
force at point x ...
and
dU1x2
Fx 1x2 = dx
Fx 1x2 = -
... is the negative
of the derivative at x
of the associated
potential-energy function.
(7.16)
This result makes sense; in regions where U1x2 changes most rapidly with x (that
is, where dU1x2>dx is large), the greatest amount of work is done during a given
displacement, and this corresponds to a large force magnitude. Also, when Fx 1x2
is in the positive x-direction, U1x2 decreases with increasing x. So Fx 1x2 and
dU1x2>dx should indeed have opposite signs. The physical meaning of Eq. (7.16)
is that a conservative force always acts to push the system toward lower potential energy.
As a check, let’s consider the function for elastic potential energy, U1x2 =
1
2
kx
. Substituting this into Eq. (7.16) yields
2
d 1 2
1 kx 2 = -kx
dx 2
which is the correct expression for the force exerted by an ideal spring (Fig. 7.22a,
next page). Similarly, for gravitational potential energy we have U1y2 = mgy;
taking care to change x to y for the choice of axis, we get Fy = -dU>dy =
-d1mgy2>dy = -mg, which is the correct expression for gravitational force
(Fig. 7.22b).
Fx 1x2 = -
221
222
Chapter 7 potential energy and energy Conservation
7.22 A conservative force is the negative derivative of the corresponding potential energy.
(a) Elastic potential energy and force as functions of x
1
U
Fx
U = kx 2
dU
2
Fx = = - kx For x 7 0, Fx 6 0;
dx
force pushes body
toward x = 0.
x
O
Potential energy is
a minimum at x = 0.
Potential energy
decreases as y
decreases.
x
For all y, Fy 6 0; force pushes
body toward decreasing y.
y
O
O
U = mgy
Fy = -
dU
= -mg
dy
an ElECTriC forCE and iTs poTEnTial EnErgy
An electrically charged particle is held at rest at the point x = 0;
a second particle with equal charge is free to move along the positive x-axis. The potential energy of the system is U1x2 = C>x,
where C is a positive constant that depends on the magnitude of
the charges. Derive an expression for the x-component of force
acting on the movable particle as a function of its position.
solution
identiFy and set uP: We are given the potential-energy func-
tion U1x2. We’ll find the corresponding force function by using
Eq. (7.16), Fx1x2 = - dU1x2>dx.
y
Solution
ExamplE 7.13
O
For x 6 0, Fx 7 0;
force pushes body
toward x = 0.
(b) Gravitational potential energy and force as functions of y
U
Fy
exeCute: The derivative of 1>x with respect to x is - 1>x 2. So for
x 7 0 the force on the movable charged particle is
Fx1x2 = -
dU1x2
dx
= -C a -
1
x
2
b =
C
x2
evaluate: The x-component of force is positive, corresponding
to a repulsion between like electric charges. Both the potential
energy and the force are very large when the particles are close
together (small x), and both get smaller as the particles move farther
apart (large x). The force pushes the movable particle toward large
positive values of x, where the potential energy is lower. (We’ll
study electric forces in detail in Chapter 21.)
Force and Potential energy in three dimensions
We can extend this analysis to three dimensions, for a particle that may move in
the x-, y-, or z-direction, or all at once, under the action of a conservative force
that has components Fx , Fy , and Fz . Each component of force may be a function
of the coordinates x, y, and z. The potential-energy function U is also a function
of all three space coordinates. The potential-energy change ∆U when the particle moves a small distance ∆x in the x-direction is again given by -Fx ∆x; it
doesn’t depend on Fy and Fz , which represent force components that are perpendicular to the displacement and do no work. So we again have the approximate
relationship
Fx = -
∆U
∆x
We determine the y- and z-components in exactly the same way:
Fy = -
∆U
∆y
Fz = -
∆U
∆z
To make these relationships exact, we take the limits ∆x S 0, ∆y S 0, and
∆z S 0 so that these ratios become derivatives. Because U may be a function
of all three coordinates, we need to remember that when we calculate each of
these derivatives, only one coordinate changes at a time. We compute the
derivative of U with respect to x by assuming that y and z are constant and only
x varies, and so on. Such a derivative is called a partial derivative. The usual
223
7.4 Force and potential energy
notation for a partial derivative is 0U>0x and so on; the symbol 0 is a modified d.
So we write
Force from potential energy: In three-dimensional motion,
the value at a given point of each component of a conservative force ...
Fx = -
∂U
∂x
Fy = -
∂U
∂y
Fz = -
∂U
∂z
(7.17)
... is the negative of the partial derivative at that point
of the associated potential-energy function.
We can
use unit vectors to write a single compact vector expression for the
S
force F:
Force from potential energy: The vector value
of a conservative force at a given point ...
S
F = −a
S
∂U
∂U
∂U n
nd +
en +
k b = −𝛁U
∂y
∂z
∂x
topography and
Potential energy Gradient The
application
greater the elevation of a hiker in Canada’s
Banff National Park, the greater the gravitational potential energy Ugrav. Think of an
x-axis that runs horizontally from west to east
and a y-axis that runs horizontally from south
to north. Then the function Ugrav1x, y2 tells us
the elevation as a function of position in the
park.
Where
the mountains have steep slopes,
S
S
F = −𝛁Ugrav has a large magnitude and
there’s a strong force pushing you along the
mountain’s surface toward a region of lower
elevation (and hence lower Ugrav). There’s zero
force along the surface of the lake, which is all
at the same elevation. Hence Ugrav is constant
at
all points
on the lake surface, and
S
S
F = −𝛁Ugrav = 0.
(7.18)
... is the negative of the gradient at that point
of the associated potential-energy function.
In Eq. (7.18) we take the partial derivative of U with respect to each coordinate,
multiply by the corresponding unit vector, and then take the vector
sum. This
S
operation is called the gradient of U and is often abbreviated as 𝛁U.
As a check, let’s substitute into Eq. (7.18) the function U = mgy for gravitational potential energy:
S
S
F = −𝛁1mgy2 = − a
01mgy2
01mgy2
01mgy2 n
nd +
ne +
k b = 1-mg2en
0x
0y
0z
This is just the familiar expression for the gravitational force.
forCE and poTEnTial EnErgy in TWo dimEnsions
A puck with coordinates x and y slides on a level, frictionless airhockey table. It is acted on by a conservative force described by
the potential-energy function
U1x, y2 = 12 k1x 2 + y22
Note that r = 2x 2 + y2 is the distance on the table surface from
the puck to the origin. Find a vector expression for the force acting
on the puck, and find an expression for the magnitude of the force.
solution
identiFy and set uP: Starting with the function U1x, y2, we
need to find the vector
components and magnitude of the corS
responding force F. We’ll use Eq. (7.18) to find the components.
The function U doesn’t depend on z, so the partial derivative
of U with respect to z is 0U>0z = 0 and the force has no
z-component. We’ll determine the magnitude F of the force by
using F = 2F x2 + F y2.
S
exeCute: The x- and y-components of F are
Fx = -
Solution
ExamplE 7.14
0U
= - kx
0x
Fy = -
0U
= -ky
0y
From Eq. (7.18), the vector expression for the force is
S
F = 1- kx2dn + 1- ky2en = - k1xdn + yen2
The magnitude of the force is
F = 21- kx22 + 1- ky22 = k2x 2 + y2 = kr
S
evaluate: Because xdn + yen is just the position vector r of the
S
S
particle, we can rewrite our result as F = - kr . This represents a
force that is opposite in direction to the particle’s position vector—
that is, a force directed toward the origin, r = 0. This is the force
that would be exerted on the puck if it were attached to one end of a
spring that obeys Hooke’s law and has a negligibly small unstretched
length compared to the other distances in the problem. (The other
end is attached to the air-hockey table at r = 0.)
To check our result, note that U = 12 kr 2. We can find the force
from this expression using Eq. (7.16) with x replaced by r:
Fr = -
dU
d
= - 1 12 kr 2 2 = - kr
dr
dr
As we found above, the force has magnitude kr; the minus sign
indicates that the force is toward the origin 1at r = 02.
224
Chapter 7 potential energy and energy Conservation
7.23 (a) A glider on an air track. The
spring exerts a force Fx = - kx. (b) The
potential-energy function.
(a)
x
-A
O
test your understandinG oF seCtion 7.4 A particle moving along
the x-axis is acted on by a conservative force Fx . At a certain point, the force is zero.
(a) Which of the following statements about the value of the potential-energy function
U1x2 at that point is correct? (i) U1x2 = 0; (ii) U1x2 7 0; (iii) U1x2 6 0; (iv) not enough
information is given to decide. (b) Which of the following statements about the value of
the derivative of U1x2 at that point is correct? (i) dU1x2>dx = 0; (ii) dU1x2>dx 7 0;
(iii) dU1x2>dx 6 0; (iv) not enough information is given to decide. ❙
A
The limits of the glider’s motion
are at x = A and x = - A.
7.5 enerGy diaGraMs
(b)
On the graph, the limits of motion are the points
where the U curve intersects the horizontal line
representing total mechanical energy E.
U
U =
1 2
kx
2
E = K + U
K
U
-A
application
O
A
x
acrobats in equilibrium
Each of these acrobats is in unstable equilibrium. The gravitational potential energy is
lower no matter which way an acrobat tips, so
if she begins to fall she will keep on falling.
Staying balanced requires the acrobats’
constant attention.
When a particle moves along a straight line under the action of a conservative
force, we can get a lot of insight into its possible motions by looking at the graph
of the potential-energy function U1x2. Figure 7.23a shows a glider with mass m
that moves along the x-axis on an air track. The spring exerts on the glider a
force with x-component Fx = -kx. Figure 7.23b is a graph of the corresponding
potential-energy function U1x2 = 12 kx 2. If the elastic force of the spring is
the only horizontal force acting on the glider, the total mechanical energy
E = K + U is constant, independent of x. A graph of E as a function of x is thus
a straight horizontal line. We use the term energy diagram for a graph like this,
which shows both the potential-energy function U1x2 and the energy of the
particle subjected to the force that corresponds to U1x2.
The vertical distance between the U and E graphs at each point represents
the difference E - U, equal to the kinetic energy K at that point. We see that
K is greatest at x = 0. It is zero at the values of x where the two graphs cross,
labeled A and -A in Fig. 7.23b. Thus the speed v is greatest at x = 0, and it
is zero at x = {A, the points of maximum possible displacement from x = 0
for a given value of the total energy E. The potential energy U can never be
greater than the total energy E; if it were, K would be negative, and that’s
impossible. The motion is a back-and-forth oscillation between the points
x = A and x = -A.
From Eq. (7.16), at each point the force Fx on the glider is equal to the negative
of the slope of the U1x2 curve: Fx = -dU>dx (see Fig. 7.22a). When the particle
is at x = 0, the slope and the force are zero, so this is an equilibrium position.
When x is positive, the slope of the U1x2 curve is positive and the force Fx is
negative, directed toward the origin. When x is negative, the slope is negative and
Fx is positive, again directed toward the origin. Such a force is called a restoring
force; when the glider is displaced to either side of x = 0, the force tends to
“restore” it back to x = 0. An analogous situation is a marble rolling around
in a round-bottomed bowl. We say that x = 0 is a point of stable equilibrium.
More generally, any minimum in a potential-energy curve is a stable equilibrium
position.
Figure 7.24a shows a hypothetical but more general potential-energy function U1x2. Figure 7.24b shows the corresponding force Fx = -dU>dx. Points x1
and x3 are stable equilibrium points. At both points, Fx is zero because the slope
of the U1x2 curve is zero. When the particle is displaced to either side, the force
pushes back toward the equilibrium point. The slope of the U1x2 curve is also
zero at points x2 and x4 , and these are also equilibrium points. But when the
particle is displaced a little to the right of either point, the slope of the U1x2
curve becomes negative, corresponding to a positive Fx that tends to push the
particle still farther from the point. When the particle is displaced a little to
the left, Fx is negative, again pushing away from equilibrium. This is analogous
to a marble rolling on the top of a bowling ball. Points x2 and x4 are called
unstable equilibrium points; any maximum in a potential-energy curve is an
unstable equilibrium position.
225
7.5 energy Diagrams
7.24 The maxima and minima of a potential-energy function U1x2 correspond to points where Fx = 0.
(a) A hypothetical potential-energy function U(x)
A particle is initially at x = x1.
U
If the total energy E 7 E3, the particle can “escape” to x 7 x4.
E3
If E = E2, the particle is trapped between xc and xd.
E2
If E = E1, the particle is trapped between xa and xb.
E1
Minimum possible energy is E0; the particle is at rest at x1.
E0
Unstable equilibrium points are maxima
in the potential-energy curve.
Stable equilibrium points are minima
in the potential-energy curve.
O
xc xa
x1
xb
x2
x3
xd
x4
x
(b) The corresponding x-component of force Fx (x) = - dU(x)>dx
Fx
O
dU>dx 6 0 dU>dx 7 0 dU>dx 6 0
Fx 6 0
Fx 7 0
Fx 7 0
x1
Caution potential energy and the direction of a conservative force The direction of the
force on a body is not determined by the sign of the potential energy U. Rather, it’s the sign
of Fx = - dU>dx that matters. The physically significant quantity is the difference in the
values of U between two points (Section 7.1), which is what the derivative Fx = - dU>dx
measures. You can always add a constant to the potential-energy function without changing the physics. ❙
If the total energy is E 1 and the particle is initially near x1 , it can move only
in the region between xa and xb determined by the intersection of the E 1 and
U graphs (Fig. 7.24a). Again, U cannot be greater than E 1 because K can’t be
negative. We speak of the particle as moving in a potential well, and xa and xb
are the turning points of the particle’s motion (since at these points, the particle
stops and reverses direction). If we increase the total energy to the level E 2 ,
the particle can move over a wider range, from xc to xd . If the total energy is
greater than E 3 , the particle can “escape” and move to indefinitely large values
of x. At the other extreme, E 0 represents the minimum total energy the system
can have.
test your understandinG oF seCtion 7.5 The curve in Fig. 7.24b has a
maximum at a point between x2 and x3 . Which statement correctly describes what happens to the particle when it is at this point? (i) The particle’s acceleration is zero. (ii) The
particle accelerates in the positive x-direction; the magnitude of the acceleration is less
than at any other point between x2 and x3 . (iii) The particle accelerates in the positive
x-direction; the magnitude of the acceleration is greater than at any other point between
x2 and x3 . (iv) The particle accelerates in the negative x-direction; the magnitude of the
acceleration is less than at any other point between x2 and x3 . (v) The particle accelerates
in the negative x-direction; the magnitude of the acceleration is greater than at any other
point between x2 and x3 . ❙
x2
dU>dx 7 0
Fx 6 0
x3
PhET: Energy Skate Park
dU>dx 6 0
Fx 7 0
x4
x
Chapter
7 Summary
Gravitational potential energy and elastic potential
energy: The work done on a particle by a constant
gravitational force can be represented as a change
in the gravitational potential energy, Ugrav = mgy.
This energy is a shared property of the particle
and the earth. A potential energy is also associated
with the elastic force Fx = - kx exerted by an ideal
spring, where x is the amount of stretch or compression. The work done by this forc
Download
Study collections